You are on page 1of 398

\e Sidéq./t: .

____________/. o~ 9~
,~ ~?
..$ ~.
t ~

ď CPSA ~ Comité Pour la Sidérurgie Ancienne


THE COMMITTE FOR ANCIENT IRONWORKING
https://independent.academia.edu/CPSA

Archeologický ústav AV ČR, Praha, v.v.i.


INSTITUTE OF ARCHAEOLOGY
OF THE CAS, PRAGUE, v.v.L
Letenská 4,11801 Prague 1 - Malá Strana; Czech Republic
http://www.arup.cas.cz

Pdf version of this boo k was created by CPSA and with consent
of the Institute of Archaeology of the Czech Academy of Science, Prague, v.v.L,
was made available on-line for those who deal with
or who are interested in early ironworking.

This pdf file is available at:


https://www.academia.edu/34485002

Reference:
Pleiner, R. (2006). Iron in Archaeology. Early European Blacksmifhs.
Praha: Archeologický ústav AVČR.
ARCHEOLOGICKÝ ÚSTAV AV ČR, PRAHA
Iron in Archaeology: Early European Blacksmiths
Iron in Archaeology
Early European Blacksmiths

RADOMÍR PLEINER

ARCHEOLOGICKÝ ÚSTAV AV ČR


Praha 2006
Published by Archeologický ústav AV ČR, Praha
Letenská 4, 118 01 Praha 1, Česká Republika

The work was realized with the support of the Grant Agency of the Czech Republic
(project No. 404/05/2063 – Iron in Archaeology: Early European Blacksmiths)

2006
c Radomı́r Pleiner

Editor for the publishers: Petr Meduna


Setting: Johana Brokešová
The English text read by: Stewart Aitchison
Cover: Radomı́r Pleiner (picture after G. Jaritz, Ferrum 77, 2005)

Printed by: Helvetica & Tempora, spol. s r.o., Praha

ISBN 80-86124-62-2

Available at: Archeologický ústav AV ČR, Praha


Letenská 4, 118 01 Praha 1, Česká republika
Fax: +420 257532288
knihovna@arup.cas.cz

Orders:
Oxbow Books, Park End Place, Oxford OX1 1HN, United Kingdom, oxbow@oxbowbooks.com
Beier&Beran - Archäologische Fachliteratur, Thomas-Müntzer-Str. 103, D-08134 Langen-
weissbach, Germany, verlag@beier-beran.de
Kubon&Sagner, Buchexport-Import, P.O.Box 341018, D-80328 Munich, Germany,
order@kubon-sagner.de
Rudolf Habelt GmbH, Am Buchenhang 1, D-53115 Bonn, Germany, info@habelt.de
Dedicated to all blacksmiths
CONTENTS

List of figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii

List of plates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .xiv

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv

Foreword and introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

I Iron in Eurasian Bronze Ages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

II The Early Iron Ages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

III Iron and steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Physical properties of iron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Recent classification of iron and steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

IV The blacksmith’s starting stock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

Bipyramidal ingots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Taleae ferreae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Ancient Greek barter iron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Roman ingots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Early medieval tool-shaped blanks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Other medieval bars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

V Forging operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

The forming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
Heat treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

VI The smith’s tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

The find complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71


Tools from smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
Smithing tools from settlements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Smithing tools from graves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Smithing tools from hoards . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
The tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Hammers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Tongs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
Cutting and piercing tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
x
Iron anvils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
Swages and dies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Files . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
Wire-drawing irons and clips . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Metal sheet shears. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .105
Forge spoons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

VII Smithing wastes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

Fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Fuel ash . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
Hammer scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
Smithing slags . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
The problem of the slag cakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Amorphous smithing slag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
Iron scrap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

VIII Smithing installations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

The hearths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123


Other installations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
The bellows and bellows protection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
The anvil as smithing equipment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
Water tanks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
Fuel stores and waste deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

IX The smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

Urban smithies of the early antiquity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135


Roman urban smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Roman rural and road smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
Smithies in Roman villae rusticae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Iron Age smithies in central and northern Europe . . . . . . . . . . . . . . . . . . . . 151
Celtic and Dacian smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
Germanic rural smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
Smithies in early medieval pre-urban centres . . . . . . . . . . . . . . . . . . . . . . . . . 160
Smithies in medieval towns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
Medieval rural smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
Medieval castle smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
Monastic smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
Smithies at mines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

X The smith’s products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

Iron inlays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184


Iron artefacts in graves and hoards . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
Iron finds from settlement layers and the scope of artefact production . . . 190
xi

XI Metallography of early irons: Reconstructed technologies . . . . . . . . . . . . 194

Simple techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196


Working of low carbon and heterogeneous wrought iron . . . . . . . . . . . . 196
All-steel artefacts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
Forge welding of carbon-poor iron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
Advanced techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
Additional carburizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
Forge welding of iron and hardenable steel . . . . . . . . . . . . . . . . . . . . . . . . . 202
Plating with steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
Steel shells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
Iron-steel-sandwich . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
Welding-in the steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
Scarf welding-on of steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
Butt-welding of steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
Top techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
Striped blades . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
Pattern-welding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
Pattern-welded swords . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
Pattern-welded lanceheads. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .219
Pattern-welded knives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
Armour making . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
Locksmithing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
Clock making . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225

XII Concise history of early European blacksmithing . . . . . . . . . . . . . . . . . . . . 226

The beginning of the use of iron and the first smiths . . . . . . . . . . . . . . . . . 226
Blacksmiths in the ancient civilizations of Greece and Rome . . . . . . . . . 228
Celtic ironworkers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
Beyond the Roman frontier: the Germanic, Cimmerian and Scythian
tribes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
Blacksmith’s work in the High Middle Ages . . . . . . . . . . . . . . . . . . . . . . . . . . 236

Glossary of technical terms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240


Glossary of historical and archaeological terms . . . . . . . . . . . . . . . . . . . . . . . . . . 244

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259

Selected abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321


Indexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
Plates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
LIST OF FIGURES

1 The earliest iron (5000 - 3000 BC). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7


2 Iron in Euroasian Early Bronze Age (3000 - 1800 BC) . . . . . . . . . . . . . . . . . . . . . 9
3 Iron ine the Middle and Late Bronze Age (1800 - 1200 BC) . . . . . . . . . . . . . . . . 11
4 Iron in the Early Iron Age in the near East and in the Late Bronze
Age in Europe (1200 - 1000 BC). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
5 Iron in the fully-fledged Iron Age (Near East) and at the beginning
of the Iron Age (Europe) 1000 - 800/700 BC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
6 The sequence of Iron Age in the Near East and Europe . . . . . . . . . . . . . . . . . . . . 17
7 The atomic arrangement in cubic lattice of iron. . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
8 Equilibrium phase diagam iron-carbide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
9 Various iron stock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
10 Neo-Assyrian bipyramidal iron bars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
11 Early bipyramidal ingots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
12 Celtic bipyramidal ingots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
13 Early ingots and bars: Schematical geograpfhical distribution . . . . . . . . . . . . . . 31
14 Manching, Bavaria. Iron ingots and sword bars . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
15 La Tène period elongated iron blanks and ‘currency bars’ . . . . . . . . . . . . . . . . . . 36
16 Iron currency in ancient Greece. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
17 Roman ingots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
18 Swedish early medieval iron blanks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
19 Scandinavian and western Slavic axe-shaped bars and accompanying ingots 47
20 Polish axe-shaped bars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
21 Medieval iron bars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
22 Forging operations: cutting, upsetting, drawing down . . . . . . . . . . . . . . . . . . . . . . 56
23 Forging operations: shouldering, bending and rolling. . . . . . . . . . . . . . . . . . . . . . . 57
24 Forging operations: splitting, twisting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
25 Forging operations: piercing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
26 Forging operations: hammer welding, swage forging, planishing . . . . . . . . . . . . 63
27 Welding-together large iron blocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
28 Comparison of the hardness of processed iron-carbon and copper-tin alloys. 68
29 Smiths and their tools on Greek painted vases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
30 Selection of Roman and Roman-style hammers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
31 Hand hammers (examples) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
32 Hammers equipped with iron shafts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
33 Sledge hammers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
34 Early set hammers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
35 Pincer tongs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
36 Hephaestus with hammer and pincer tons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
37 Tongs with pointed ans S-shaped universal jaws . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
38 Smithing tools as symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
39 Tongs with adapted jaws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
40 Smith’s chisels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
41 Punches and drifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
42 Iron block anvils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
xiii
43 Block anvils with pointed bases as in Roman reliefs . . . . . . . . . . . . . . . . . . . . . . . . 95
44 Sens, France. A Roman tombstone of a smith with
hammer and anvil with nail-hole . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
45 Horned iron anvils. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
46 Field anvils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
47 Vlastislav hillfort, Bohemia. A set of blacksmith’s tools . . . . . . . . . . . . . . . . . . . . 100
48 Dies and swages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
49 Nail- and/or wire-drawing irons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
50 Files . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
51 Metal sheet shears . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
52a Forge spoons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
52b Bygland, Norway. A rectangular forge spoon from a hoard . . . . . . . . . . . . . . . . . 107
53 Hammer scale from a Roman smithy at Nailly, France . . . . . . . . . . . . . . . . . . . . . 111
54 Smithing slag cakes (plano-convex hearh bottoms, PCB) . . . . . . . . . . . . . . . . . . . 121
55 Installations in smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
56 Pre-Roman smithies in Europe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
57 Roman smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
58 Blagaj-Maslovare, Bosnia. Roman smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
59 Smithies in Roman villae rusticae and in the Romano-Barbarian Snorup . . . 159
60 Smithies in medieval settlements and suburbs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
61 Medieval rural smithies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
62 Lelekovice, Moravia, castle smithy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
63 Smithies at mines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
64 Iron inlays on bronze artefacts. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
65 Býčı́ Skála cave, Boravia. Bull statuette . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
66 Development of the assortment of iron artefacts in ancient Greece . . . . . . . . . 191
67 Drilling with early medieval augers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
68 Reconstruction of the manufacture of an Avar sabre . . . . . . . . . . . . . . . . . . . . . . . 207
69 Technology of the manufacture of early Slavic battle axes . . . . . . . . . . . . . . . . . . 209
70 Axe and adze . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
71 Technology of the manufacture of early medieval knives. . . . . . . . . . . . . . . . . . . . 213
72 Early La Tène period pattern-welded sword . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
73 Two basic systems in the application of pattern-welded blades . . . . . . . . . . . . . 217
74 Examples of early medieval pattern-welded swords . . . . . . . . . . . . . . . . . . . . . . . . . 218
75 Examples of display lanceheads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
76 Reconstruction of the manufacture of an early medieval pattern-welded knife 221
LIST OF PLATES

Pl. I. Metallographic structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341


Pl. II. All-iron sword blade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
Pl. III. All-iron weapon and tilling tool . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
Pl. IV. Application of mild steel in tool manufacture . . . . . . . . . . . . . . . . . . . 344
Pl. V. Forming, folding and welding of structural iron . . . . . . . . . . . . . . . . . 345
Pl. VI. Early all-steel blade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
Pl. VII. All-steel and addtionally carburized artefacts . . . . . . . . . . . . . . . . . . . 347
Pl. VIII. Piling of mild steel bands. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
Pl. IX. Piling and folding of heterogeneously carburized iron . . . . . . . . . . . 349
Pl. X. Piling of carbon-poor iron in toolmaking . . . . . . . . . . . . . . . . . . . . . . . 350
Pl. XI. Secondary carburization of blades . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
Pl. XII. Piling of carburized bands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 352
Pl. XIII. Secondary carburization of a knife blade . . . . . . . . . . . . . . . . . . . . . . . . 353
Pl. XIV. Iron and steel plating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 354
Pl. XV. Welding-on steel as applied on tillage implements . . . . . . . . . . . . . . . 355
Pl. XVI. Welded-on steel shell on a weapon blade . . . . . . . . . . . . . . . . . . . . . . . . 356
Pl. XVII. Welding-on of phosphorus-rich shells . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
Pl. XVIII. Application of the three-layer ‘sandwich’ in cutlery . . . . . . . . . . . . . 358
Pl. XIX. Application of the iron-and-steel three-layer system . . . . . . . . . . . . . 359
Pl. XX. Welding-in an iron-and-steel pile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
Pl. XXI. Inserting steel cutting-edge into axeheads and knives . . . . . . . . . . . 361
Pl. XXII. Folded blade inserted into a socket . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362
Pl. XXIII. Inserting a steel point into a medieval mining pick . . . . . . . . . . . . . . 363
Pl. XXIV. Scarf welding-on of a steel cutting-edge on an axehead . . . . . . . . . . 364
Pl. XXV. Scarf welding-on of steel in cutlery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
Pl. XXVI. Scarf welding-on of a caburized cutting-edge . . . . . . . . . . . . . . . . . . . . 366
Pl. XXVII. Scarf- and butt-welding of steel in cutlery . . . . . . . . . . . . . . . . . . . . . . 367
Pl. XXVIII. Butt-welding of steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368
Pl. XXIX. Heat treatment applied on a wire-drawing iron . . . . . . . . . . . . . . . . . 369
Pl. XXX. Butt-welding of steel cutting-edges in cutlery . . . . . . . . . . . . . . . . . . . 30
Pl. XXXI. Striped damast as applied on medieval knives . . . . . . . . . . . . . . . . . . 371
Pl. XXXII. Pattern-welded panels in a sword blade from Bešeňov. . . . . . . . . . . 372
Pl. XXXIII. Bešeňov (continued) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
Pl. XXXIV. Manufacture of an early medieval princely sword . . . . . . . . . . . . . . . 374
Pl. XXXV. Pattern-welding on weapons and knives. . . . . . . . . . . . . . . . . . . . . . . . . 375
Pl. XXXVI. Manufacture of an early medieval display lancehead . . . . . . . . . . . . 376
Pl. XXXVII. Pattern-welded knives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377
Pl. XXXVIII. Pattern-welded cutlery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378
Pl. XXXIX. Steel in the manufacture of medieval plate armour . . . . . . . . . . . . . . 379
PREFACE

It is more than 50 years since Radomıı Pleiner published his first paper on ancient iron
working. Since then he has produced more tha 250 scientigic papers and several books on
this important topic. He has been, without question, one of the leading scholars in this
field.
The establishment of the metallurgical laboratory in 1963 at the Prague Institute of
Archaeology gave Pleiner a permanent base for his work. In 1966 he founded the Comité
pour la sid0rurgien ancienne (CPSA) and remained its secretary for nearly 40 years.
During this period he has played a vital role in the rapid development of the discipline, by
collating new research, publication and information on work in progress, and by publishing
abstracts in the CPSA Communications printed twice yearly in Archeologické Rozhledy.
These abstracts conprise several thousand items - a crucial source of information in the
pre-internet age. Important factors in this have been Pleiner’s linguistic skills, enabling
him to act as a bridge between the (old) east and the west with his enomrmous energy.
The contacts he has made and encouraged through this work have rsulted in him being a
guiding light for several generations of iron scholars.
No one has used this wide ranging source better than Pleiner himself, as witnessed by
his magisterial volume on ‘Iron in Aechaeology: The European Bloomery smelters, and
now by this companion volume which synthesises the evidence for secondary iron working
and smithing. This volume must have been significantly more difficult to produce, partly
because of the fragmentary and very varied nature of the archaeological material, but
mainly because of the loss of Pleiner’s archive of samples and notes, and the library of
the Institute of Archaeology, during the disastrous floods of 2002. This would destroy the
spirit of a lesser man, but the fact that Pleiner has been able to complete this volume with
little delay is tribute to hid determination an to his thorough knowledge of the source
material.
Future generations of iron working scholars will long remain in his debt.

Peter Crew
xvi
FOREWORD AND INTRODUCTION

Some time ago my book entitled ‘Iron in Archaeology: The European Bloomery Smelters’
was published (Prague 2000). Originally, according to a project by the British archaeomet-
allurgist Henry Cleere (to which I was kindly invited to contribute), it should involve both
branches of the ancient iron technology: the smelting and working of the metal. It turned
out that it was impossible to achieve this it in one volume so that, after Dr. Cleere was
fully engaged at the UNESCO, I continued with his permission and help to compile and
finish the first volume. This presents a survey of all sources concerned with the ancient
and medieval history of the direct iron making process - the bloomery work. The themes
dealing with the early development of the iron industry include production regions, raw
material exploitation, principles of the bloomery process, smelting and reheating installa-
tions and the products: the blooms. The second wide field of the technology of iron, the
further working of the metal, remained untouched. This, as expressed in the final phrase
of the book’s introduction, has had to be reserved for a specific volume. Keeping this
promise I have compiled the following pages and fulfilled Henry Cleere’s initial idea.
In carrying out my task I could take an advantage of an experience: in 1962 appeared
my book ‘Staré evropské kovářstvı́ - Alteuropäisches Schmiedehandwerk’ (Prague) which
took under discussion the early European blacksmith’s work until about AD 1200. This
was inspired with the presentation of the first set of 63 metallographically investigated iron
objects from the territory of the former Czechoslovakia (produced by the author in the
newly established metallographic laboratory in the Archaeological Institute at Prague).
The results were compared with what had been published until that time in European
countries, most if which came from Poland (Jerzy Piaskowski) and Russia (Boris A.
Kolchin). At the same time information was gathered on various sources and materials
concerning the craft. However, after more than forty years the source base widened in
such a manner that a new approach has had to be applied and, simultaneously, a thorough
selection of data carried out.

The present volume is divided into twelve chapters of differing extent. The first part of
the topic deals anew with the discovery and spread of ironworking in the Old World. The
subsequent parts discuss the working and utility properties of early iron and steel and the
principal sources yielded by archaeology: the blacksmith’s stock (bars, ingots, scrap), his
waste products (especially smithing slag), his installations and workshops – the smithies.
then, in a very abridged form, the actual products of the smith are described: iron objects
as they appear in graves, hoards and settlements.
Chapter XI is devoted to the metallography of early iron artefacts which offers im-
portant data on their utility properties and above all on their construction: simply and
sophisticatedly forged items. In other words, it reveals the ‘anatomy’ of forgings. Met-
allography distinguishes crystal structures of metals and their behaviour under different
heating and working conditions. From modest beginnings in early years of the 20th cen-
tury the examinations developed to systematic study which disposes of about 14 000
metallographic analyses of iron forgings from different European countries have been
published. Generations of metallographers and archaeometallurgists have been engaged
in this research. The work is going on. Certain results achieved by the metallographic
2 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
examination of early artefacts have been simulated by experimental forging during which
the effects of the piling, welding together soft iron and hard steel, shifting carburized
zones under hammer strokes, twisted pattern-welding etc. could be studied.

A concise history of early European blacksmithing as reflected in studied sources con-


stitutes a concluding chapter.

I have to confess that my work has suffered. Serious difficulties arose after the catastrophic
floods in August 2002. The water in my basement laboratory and office reached the
ceiling and all books, investigation reports, notes, adresses, slides and photographs were
destroyed except those with which I had been working at home before the disaster. What
is worse, the library of the Archaeological Institute which was situated in the basement
and has gone, too. I am gratefull to many of my colleagues, at home and abroad (and
to corresponding members of the Comité pour la sidérurgie ancienne which whom i have
been in contact) for their help and for providing me with some reprints, offprints and
copies. To completely acquire the lost materials is naturally impossible and the revision
of data is also frequently problematic.
Nonetheless, in spite of these conditions I tried to finish the book in an appropriate
manner. The goal is to present a guide for all interested readers which could provide them
with basic information on the topic in question, its sources, evaluations and bibliography.
I am adding two glossaries: one explains some technical, the other some historical notions
which could not be described in detail in the text.
I hope that the book proves useful for those who are prepared to continue in the research
in the scientific field of the history of the working of iron in earlier human history.
Chapter I

IRON IN THE EUROASIAN BRONZE AGES

The discovery of iron and the increasing frequency of iron artefacts took place in periods
which in archaeology are denoted as Bronze Ages when copper-based alloys dominated
the manufacture of tools, weapons, and ornaments. The reason for using the plural form
is inherent in the fact that the process of spreading iron differed in the late period of the
Bronze Age in chronological eras and in individual parts of Eurasia. The classifications
of the Bronze Age are expressed by various notions like e.g. Hittite, Cypriot, Cypro-
geometric, Minoan, Helladic in the eastern Mediterranean and the Middle East where
the Schaeffer’s high and Piotrovski’s low chronologies can be applied (Schaeffer 1948;
Piotrovskiy 1948).
In Europe, the chronologies by Paul Reinecke and Hermann Müller-Karpe cover the
Bronze Age from the beginning of the 2nd millennium BC up to the Hallstatt B3 period
(9th - 8th centuries BC) which is marked by the sporadical occurence of iron artefacts
in the central part of the Continent. For Scandinavia, Oscar Montelius elaborated a
chronological system which divided the Bronze Age into six periods; sections V and VI
correspond with the H B to H D in central Europe where, at that time, the early Iron
Age began. In all of the regions the late Bronze Age witnessed rare and gradually more
frequent appearance of iron, i.e. artefacts and their fragments, or metallurgical waste
(slag), and even blooms.
It is presupposed that iron was recognized as a byproduct of copper smelting using
chalcopyrite or charges of copper and iron ore - the latter having been added as a flux
for easier melting of copper slag (e.g. Charles 1980; for an opposing and unncessarily
general wiev see Merkel and Barrett 2000). Under favourable metallurgical conditions
metallic iron could be found embedded either in the copper slag (which was fayalitic
and wüstitic with a minimal copper content) or even in copper cakes and ingots. After
long periods of smelters’ experience this metallic iron was discovered as a new metal,
separated and used for making small precious artefacts. This slow process took place in
ancient regions with a very long tradition of copper production. Nonetheless, the question
is where it appeared for the first time, and when: i.e. in what period iron appeared as
a specific metal with different properties in comparison to copper and in what period it
began to be produced intentionally. The basic problem is the chronology - the dating
of artefacts, wastes, installations, and the dating methods. All kinds of archaeological
sources should be dated according to one method to get a chronological sequence. The iron
artefacts in question have been dated according to an archaeological system of chronology,
especially those from graves and hoards. These systems are based on the sequence of
historical events derived from written sources and their absolute position in time. On
the other hand, metallurgical wastes (slags) often suffer from the lack of accompanying
archaeological dating material and are frequently dated using physical methods (mostly
radiocarbon or archaeomagnetic measurements) which produce absolute data: in spite of
some agreement the combination of dating methods reveal serious discrepances in many
cases. The radiocarbon data regularly show earlier figures than the archaeological systems
and provide wide ranges of time spans. Moreover, these datings can result in a certain
amount of obscurity. An example is provided by the earliest ironworks grouped in the
4 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
ancient Colchis where the radiocarbon measurements and archaeological sources (pottery)
cover a time span of 1200 years (see below). Many scholars in different parts of the Old
World like to see metallurgical and other events taking place as early as possible and
prefer the higher dates.
Another question is the spread of iron from certain geographical focuses: the circulation
of small ready-made artefacts (ornaments, amulets, even daggers) as export/import, i.e.
spoils, gifts among prominent people (both known from written sources), initially rarely
in the form of trade. The other possibility is the spread of knowledge - this is a very
complex problem because, under given conditions, the know-how of the use and working
of iron was integrally bound to individuals, the metalworkers, who must have migrated
either together with e.g. raiding and expanding groups or even in the opposite direction of
such a streaming. In new places they worked, in early periods, on behalf of their masters
(craftsmen ‘of the court’). They became assimilated and could produce iron artfacts if
they had some material at their disposal, such as copies or imitations of common bronze
objects. The third way of distribution and spread was the export/import of certain
quantities of unworked metal: again spoils, ‘royal gifts’, later tributes and taxes, and
finally long-distance trade. This phase exceeds the frame of the archaeological Bronze
Age. Communication among metalworkers has to be presupposed as a result of deep
changes in social organization in many cultural provinces.
Was the metal iron identified, invented and spread from one region or was the secret of
its smelting discovered in more than one place? Monophyletic and polyphyletic approaches
come into question here. Many archaeologists and archaeometallurgists would like to
see evidence for independent development in the territories of their interest. This has
recently become a serious theme when considering the efforts of e.g. African scholars (see
below). However, the convergency of technological discoveries in various focuses (fields
like agriculture, pottery, textiles) is not so easily acceptable in the case of metallurgy
(especially that of copper and iron) which presupposes an extremely long development
in an environment with specific conditions - not only accessible ore and fuel resources
but above all social background. The possibilities for the inception of such a process in a
number of areas were considerably restricted. If it should be admitted that any innovative
technology could be transferred in the direction of powerful streams of cultural influence,
than it can be hardly expected that the same technology could be independetly invented
in the path of such pressure.
Leaving aside the dating and spread we must consider another factor: the fragmen-
tariness of the sources, both archaeological objects and written records. The first are ex-
tremely heterogeneously distributed in space and their abundance depends to a great deal
on the intensity of archaeological research. The latter are problematic, moreover, when
dealing with specific facts: their authors might have provided distorted and erroneous
data the interpretation of which rests bases on the correct reading and understanding of
the terminology (and they should be reliably dated as well).
Despite the fact that very early specimens of iron have been sporadically found in the
Near East dating from the 5th millennium BC (Waldbaum 1980, 69 - 70), the evolution
of the nomenclature of iron first appear in the late 3rd millennium BC (an.na, kú.an,
amūtum, aššium). Recently, the terminology has been submitted to criticism (the terms
express various meanings and some readings of cuneiform texts leave these words un-
translated, see Richter 1997). The unambiguous identification of the new metal is offered
by texts dating from the early 2nd millennium BC onwards (AN.BAR, parzillu, hapalki).
IRON IN THE EUROASIAN BRONZE AGES 5
What is not pleasant is that in regions where written sources appear, a discrepancy in
the frequency of archaeological iron and written references may occur. There are regions
with relatively abundant archaeological relics in question and others with a minimum of
archaeological objects but which are elucidated by the testimony of historical texts.
The written sources enable to consider the use of iron in another light: weigth. The
precious and rare iron objects from the Near East up to the mid-second millennium BC
weighed grammes or shekels (about 16g each); later, in the advanced Hittite period (1500
- 1200 BC) minae, MA.NA are mentioned (about one pound each, see Pleiner 1996a). At
the time about 800 BC central Europe saw iron daggers, sword blades and even blooms
comparable with the minae economy but the North of Europe used iron in shekel amounts
at that time and 2 - 3 centuries later still. The written sources from the decline of the
Hittite period in Anatolia mention bars of iron (PAD AN.BAR) weighing 1.5 - 2 minae
in quantities of 14 to 38 units, i.e. 17 to 19kg (KBo XVIII. 155, see Siegelová 1984, 157).
The subsequent centuries witnessed an enormous rise of iron production which is also
reflected in the weigth figures (see the section devoted to the early Iron Age below).

The use of meteoritic iron in early history is the subject of a discussion which cannot
be neglected (see also Chapter III) . Two points of view have been presented. The first
regards as realistic the possibility that fragments of sideric meteorites, mostly octahedrites,
containing nickel exceeding several mass percents (4 - 10% and up to 19%, plus about
0.5% Co and 0.1 - 0.9% P) were occasionally used for manufacture of extremely precious
objects. Previously published bulk analyses (see e.g Waldbaum 1980; Tylecote 1987, 97 -
105) which gave higher nickel contents cannot be checked any more. However, the nickel
is distributed unevenly in the metal and since the results represent average values they
indicate considerable enrichment in the main. Meteoritic iron is malleable both in its hot
and cold state (Tylecote o. c.). The Inuit Eskimos exploited the Cape York meteorite
in Greenland in modern times (Craddock 1992; Buchwald 2002). A certain number of
very ancient iron objects analysed in the early 20th century revealed 3 - 11% Ni and
have usually been declared as made of meteoritic iron (for surveys see Waldbaum 1980;
1999, both with references). Support is provived by the early nomenclature of iron as
the ‘metal of heaven’ (Egyptian bi’a.npet, Hittite AN.BAR.GE6 , see Siegelová 1984, 159 -
163) and the fact that falls of meteorites are mentioned in the Near East cuneiform texts
(Bjorkman 1973). Meteoritic iron, when available in small quantities, involved, without
any doubt, specific qualities appreciated in ritual contexts and could be have been used
for making artefacts for a special purpose (Buchwald 2005, 13 - 38).
This traditional view was subjected to criticism in the 1980’s when it turned out that
some nickel-rich iron ores like the Greek laterites migh have yielded nickel-containing
iron (Photos 1989; however, Tylecote 1987, 99 says that the Greek laterite ores could
produce iron with mere 1 - 2% Ni); possibly the same effect could have been produced
by charged admixtures of chloanthite (nickel arsenide NiAs) as Piaskowski has proposed
(1980). Moreover, modern microanalyses have shown that the nickel, when present in iron,
segregated in considerable amounts in the welding seams of artefacts including medieval
items.
Nevertheless, modern research has brought evidence of the use of meteorites in the
manufacture of exclusive objects (Bichkin Buluk and Boldyrevo in Transcaucasia, the
latter producing an iron blade in copper socket (9.1 - 9.45 % Ni, 18th century BC, see
6 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Terekhova in Ocherki 197, 33 - 39; Pleiner 2000c, 27. fig. 1: 10); In China bi-metallic
objects with nickel-rich parts have been recognized as well: Kung Chien in Hopei, Li
Chung 1979; axes from museum collections, see Gettens et al. 1971.

Despite the above problematic points pondering on the following phenomena allows the
presentation of a reasonable sequence of events regarding chronology. The first is the
quantity of iron that was available in the early days of the late Bronze Age. It can
be estimated not according to incomplete numbers of iron objects unearthed but rather
by considering their size and weigth: there are items comprising a minimal amount of
material - iron inlays, amulets, ornaments, knives and razors which consumed grammes
or tens of grammes of metal (the shekel economy as mentioned above). Later, certain
regions saw pounds or kilograms expended to make other kinds of objects: sword blades,
axeheads, chisels, bars, ingots and blooms. In Anatolia and other parts of the Near East
we encounter them during the 2nd millennium BC (see below) but centuries later rare
blooms were found in complexes of the final Bronze Age in Europe (8th century BC)
as at Šafárikovo-Tornala in southern Slovakia weighing 2.41kg - 5 minae (Pleiner 1981,
121; Furmánek 1988, 187, fig. 3) and the Romanian Sincraieni (weight not given, see
Boroffka 1991, 11). The bloomeries that produced them are not known but the finds of
that category of weight indicate the spreading or the iron technology in this case in a
north-westerly direction from the assumed epicentre.
This epicentre must have been somewhere in the eastern Anatolia where the metallur-
gists recognized a new metal with entirely different properties.
Here, it is unavoidable to take into account the mythological tradition and historical and
archaeological facts concerning Asia Minor. First, the legend of Phrygian Dactyli cannot
be omitted. Dactyli (Daktyloi) were mythical dwarfs (fingers or thumbs) originating in
the Mount Ida in Phrygia, Asia Minor, who, according to the tradition, were the first
who produced iron (Diodorus V, 64 - 65; Strabo 10.3.22, Marmor Parium). Already in
the antiquity the Phrygian Ida used to be confused with Mt. Ida, Crete. Dactyli were
in service of Rhea, the Mother of Gods as well as other beings: Curetes, Corybantes,
Cabiri, Tibareni, Telchines which are considered to have participated in the beginnings
of metallurgy. Presumably they were priests of Rhea. Seyffert (1902, 316) mentions
three Dactyli by names: Celmis (allegedly the smelter), Damnameneus (the hammer) and
Acmon (the anvil). The myth of bronze- and ironworking Dactyli appears in works of
other classical writers as well. It seems to be very ancient but it belongs to legends ,
whilst the other tradition conerning the earliest iron smelting, more realistic, is that of
the ironmaking Chalybes (sidérotektones by Aeschylus), in the NE Anatolia. As a matter
of fact, these traditions do not contradict the facts.
The art of smelting and working it spread to all directions. During the 8th century BC
it even reached the North-West of China (Wagner 1997).
The earliest irons (14 in number) come from the Middle East and Egypt (Waldbaum
1980, 69 - 71). The 5th millennium piece from grave A at Samarra, Iraq, was 4.3cm long
and was apparently smelted, the other four objects are said to be of meteoritic origin due
to the Widmannstätten structure and high nickel content (one of the Egyptian beads from
El Gerzeh: 7.5% Ni). All of these objects would be classified in the category of grams or,
as mentioned, shekels in terms of the Near Eastern metrical system. At that time, the
surrounding territories lived in Neolithic or Chalcolitic civilizations (Fig. 1).
IRON IN THE EUROASIAN BRONZE AGES 7
8 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
The period from about 3000 to 1800 BC produced nickel-bearing artefacts in both
regions mentioned (6 items) plus a further three from 18th century BC burials north of
the Caucasus (see above). Smelted nickel-free artefacts (6) appeared in the Near East as
well. The remainder (10) have not been analysed. Apart from small objects and fragments
first of all weapons have to be registered (e.g the display dagger from grave A at Alaca, the
sword from Dorak, the so-called macehead from Troy, all in Anatolia, or a dagger with an
iron blade from Tell Asmar, Mesopotamia). They might have reached the minae weights,
about 1 pound each. Iron artefacts have regularly been recovered from rich if not princely
graves which indicates that the metal was reserved for the upper class of society. Special
attention has to be paid to a stratified piece of white cast iron (3.5% C) from Geoy Tepe
at lake Urmia in northern Iran (Burton Brown 1950), presumably accidentaly produced
and discarded as unworkable waste. Experimental smelting activities took place in that
part of the world. Cuneiform texts of the centuries around 2000 BC, connected with rites,
oaths and other deeds of prominent individuals contain terms which use to be explained
as denoting iron (see above) and in the early 2nd millennium BC the AN.BAR, parzillu,
barzel definitely ment iron. (Fig. 2). The period between 1800 and 1200 BC seems to
have been the most critical in terms of the beginnings of iron making and working in the
Near East and Greece (Fig. 3). The number of artefacts discovered in Anatolia reaches
the order of 50 and the cuneiform late Hittite records are aware of smelted and meteoritic
iron and include the first report on intentionally (but with difficulty) smelted iron for
daggers (the Hattushil KBo 14 letter). The start of certain radiation can be observed: the
Aegean with Greece (more than 20 objects, iron slags around 1200 BC), Egypt (imports
from the Mitanni realm (the eastern neighbour of the Hittite confederation) attested by
written records), the area between the Dniepr and Don rivers in Ukraine and southern
Russia (weapons, poorly described hearths from Voronesh (Shramko 1981). Isolated finds
emerged in the Balkans and sub-Carpathian regions (Boroffka 1991; Furmánek 1988) as
well as in Sicily, Sardinia and even central Europe (Kimmig 1964, 274 - 283; Pleiner 2000c,
23 - 33, with references). At Baageroostwelde in the Netherlands a peculiar iron awl was
found which was dendrochronologically dated to the mid-14th century BC, see Charles
1984). Earlier is solely the iron dagger hilt from the ritual well of the Otomani culture
at Gánovce in northern Slovakia, 15th century BC (e.g. Pleiner 1981, 115, fig. 1, with
reference). The validity of the above hypothesis about the spreading of this still precious
and prestigious commodity seems to be highly probable if not self-evident (Fig. 3).
A critical approach to all the mentioned sources concerning the earliest history of iron
has to be taken in the subsequent era in the development of human culture - the Early
Iron Age but, in fact, no objections can be brought against the presented theory.
IRON IN THE EUROASIAN BRONZE AGES 9
Chapter II

THE EARLY IRON AGES

Archaeological classifications do not use always this term. Phases are also named
according to important sites: Hama in Syria, Mycenae in the NE Mediterranean, Hallstatt
and La Tène in Europe, or after ethnic groups like Cimmerian and Scythian or Etruscan
or even ornamental style (Protogeometric and Geometric in Greece). The chronological
sequence in various parts of Eurasia is demonstrated in Fig. 6 based on works by Wright,
Waldbaum, Snodgrass, Shramko, Reinecke, Müller-Karpe, Alexander, Scott and Nørbach.
The archaeological evidence, i. e. the slowly increasing frequency of iron objects is
made up of various indicators, the number of artefacts uncovered by archaeology (the less
reliable factor), the kinds of artefacts (weapons, tools, ornaments, fittings, utensils etc.),
the weight category (as discussed in the previous chapter), the presence of production in-
stallations and metallurgical waste. As for Anatolia, Syria and the eastern Mediterranean
of the 12th - 10th centuries BC the numbers of iron objects vary greatly but in some places
exceed 200 to 300 items (Syria, Palestine, Greece with Crete and the islands but merely
16 in Anatolia). An increase of kinds of objects can be observed, more than thirty: knives
and ornaments, also armour scales, chisels, nails, rivets and agricultural implements have
to be added to weapons as well. Metallurgical wastes, probably smithing slags were,
announced suo tempore from Vardaroftsa in Macedonia (Davies 1926/27). Fig. 4
Of a specific importance are, however, the discoveries in Georgia, ancient Colchis, a
territory adjacent to the eastern coast of the Black sea having yielded what are still unique
concentrations of ironworks. The sites were excavated by Gzelishvili (1964) and above
all by Khakhutaishvili (1987). Four concentrations with several grouping of ironworks
were registered, nearly 200 workshops alltogeher. Metallurgy is not discussed here (for a
survey see Pleiner 2000c, 36 - 38). Instead, the chronology and economic influences are
considered in a brief note. Dating problems which have already alluded to above should
be mentioned first. The pottery finds indicate the period of the 9th to 7th centuries BC.
Radiocarbon datings and their calibration show, in certain cases, much higher figures.
E.g. the site of Mziani I dates from about 720 BC, that of Mshvidobauri II 1810 BC
(archaeomagnetic measurements 1207±100BC). What is worse, in several cases one and
the same furnace yielded different data: from furnace Mziani III 1, one charcoal sample
was dated to 640 BC, the other taken from 10cm below to 1020 BC. Furnace Mziani
II produced different charcoal dates from various of its levels: 575/628BC, 20cm lower
940/1072 BC, and a sample from the very bottom 1280/1495 BC (Khakhutaishvili 1987,
128, 136, 150, 179-180). It should be born in mind that a smelt performed in such a kind
of pit furnace could have only lasted for several hours at the most. Moreover, the layout
of the excavated bloomeries in all of the groupings reveals an identical character: one or
two pit furnaces surrounded by or adjacent to a slag heap or layer (20 to 100m square,
30cm to 100cm thick, volumes 4 to 6m in volume thrown slope downwards) and burnt-red
and stone-paved place behind which apparently served for reheating purposes. It would
be striking if this pattern were to have been applied for more than 1200 years. In the light
of these facts one would be inclined to believe in the archaeological dating as proposed
for the ironworks at Djikhandjuri, Legva, Charnali, Choga (8th/9th - 7th centuries BC)
or archaeomagnetic measurements (570± to 1026± BC). At any rate, the concentrations
THE EARLY IRON AGES 11
12 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
of ironworks of the ancient sub-Caucasian Colchis, to the north-east of the legendary
Chalybes, are so far the oldest known and investigated. No estimation concerning the
yield of ironmaking in Colchis is presented in the cited texts but using parameters of slag
amounts given by Bachmann (1982, 4-5) one Colchidian site could have produced about a
hundred tonnes of unworked iron during its existence which would mean that the region
could have delivered some two thousand tonnes in about four centuries. If we calculate
that half disappeared during the reheating and smithing the sum would be equivalent to
more than thirty thousand biltu - talents (cf. Pleiner 1996a).
Where were the consumers? A partial answer is offered by the royal inscriptions in Neo-
Assyrian centres in modern Iraq. They provide information on iron which was received as
spoils and subsequently as tributes and taxes from various parts of the Near East (Pleiner
and Bjorkman 1974, 291 - 294, with references). It is remarkable how these amounts
increased during the 9th and 8th centuries BC since the times of Tukulti-Ninurta (890 -
884 BC) up to Tiglath-Pileser III (744 - 722 BC): they are given in talents (biltu) from
1 or 2 up to 90 000. A part of the prescribed taxes might have even originated in the
Colchis because places in the north are already named in the 9th century: Nairi, Zamani,
Shupria, Tushpa in Urartu delivering 100 - 300 talents. But the bulk of the neigbouring
city states and regions the number of which increased to about 25 in the late 8th century
(o.c. figs 6 and 7) were more in the south in eastern Anatolia and northern Syria (e.g.
Que, Hattina, Gorgum, Carchemish charged with 250 talents, Halupe etc.) and later
in Palestine as well. A huge booty came from Damascus, mentioned on the Nimrud and
Rimah steles - 5000 or 2000 talents, i.e. about 150 tonnes of iron. Tiglath Pileser collected
probably the largest amount of iron - a literary text known as the ‘Letter to Gilgamesh’
requests an incredible quantity of goods, among them 90 000 talents (3 000 tonnes) of
‘high quality’ iron. Maybe that the Assyrian rulers boasted about the real sums and
exaggerated the figures. Nonetheless the store of Sargon II at Khorsabad contained 16
tonnes (5 300 talents) of iron in the form of bipyramidal bars (see Chapter II 2), chains
and rings (Place 1867/1870; see Pleiner and Bjorkman, o.c., 245, fig. 8).
Thus, the iron making region in Colchis was not the only producer and the small city or
princely states in eastern Anatolia, Syria and Palestine were able to deliver tonnes of the
black metal. Where were their bloomeries situated? One bloom weighing 2.5kg represents
about 10kg of bloomery slag, one talent about 100kg. Twenty eight thousand tonnes of
worked iron delivered to Assyria during the 9th and 8th centuries BC must have left not
85 000 tonnes of slag but twice as much when reheating and working losses are taken into
account. Unfortunately, no further ironworks have been discovered in Anatolia, Syria and
Palestine which is possibly due to research strategy which has been orientated towards
urban settlements and temples.
At that time the Near East and eastern Mediterranean adopted a fully-fledged iron-
based civilization. In other parts of Europe what should be called the Initial or Beginning
Iron Age began in a stream of radiation. The Balkans might have entered this stage during
the 8th - 7th centuries BC but in central and western Europe the situation changed in the
subsequent Hallstatt C and D and La Tène periods, during the 7th to 5th centuries BC.
These cultures witnessed the penetration of iron technologies but still on a limited scale.
Few ironworks were uncovered (a survey see Pleiner 2000c, 32, 58) and more than twenty
kinds of iron artefacts were produced: apart of ornaments and weapons like flange-hilted
swords, daggers, spear- or lanceheads and knives - tools were forged: smithing tools (see
Chapter VI), chisels, sickles. Of importance is that horse gear was in service of prominent
THE EARLY IRON AGES 13
14 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
members of society and nailed iron wheel tyres were used in the construction of their
carts. Iron fire-dogs of the low long type appeared among archaeological finds. About
half of these groups of objects may be included into the shekel and the rest into the minae
categories. In Scandinavia, the process was slower. Apart from some metalworks which
produced some iron and the dating of which is the subject of discussion (Hjäthner-Holdar
1993, 62 - 99; Zimmermann 1998; see Pleiner 2000c, 25) the inventory of iron things
remained more meagre: about fifteen kinds, mostly inlays, ornaments and their parts,
rarely Hallstatt type swords, spearheads, knives and chisels, Hjärther–Holdar o.c. 121 -
185). Most of the items are of shekel weights (Fig. 5).
The Middle and Late La Tène periods of central and western Europe signalize, at least
in the areas settled by Celts, the coming of the fully-developed Iron Age has yielded many
sources concerning individual aspects of ironworking. They are the subject of further
explanation and discussion in subsequent chapers.
At the last it seems to useful to summarize the characteristics of the Iron Ages as to the
functional role of iron, especially in Europe. The first stage may be called pre-Iron Age:
small iron amulets, ornaments and ceremonial and symbolical weapons of rulers (in the
European Greek world Late Minoan and Late Helladic periods have to be mentioned; just
sporadical occurence of minute iron samples, e. g. in hoards of bronzes can be observed
in central Europe of the Hallstatt A period which represents the full Bronze Age). Iron
has been recognized presumably as a by-product of the copper production.
The next phase would be the proto-Iron Age: Artefacts made of precious iron occured
in tens of cases. In Greece this concerns the Protogeometric and partly Geometric period,
in central Europe the Late Bronze Age Hallstatt B3 period (9th/10 centuries BC). Iron
knives, small iron objects and rarely even iron blooms of unknown origin appeared among
archaeological metal finds. However, the upper class of the society still controlled the
production and handling with iron, at that time produced intentionally from iron ores.
The Early Iron Age can be divided into two subphases: the Initial and the Developed.
The first saw the use of iron weapons, ornaments and first tools - except knives first
iron axeheads, chisels and sickles eased the working processes within the Hallstatt C/D
and La Tène A periods (7th/6th to 5th/4th centuries BC). At that time, the eastern
Mediterranean entered the fully-fledged Iron Age. During the Developed phase of the
Iron Early Iron Age the technical black metal spred to wider social groups. Warriors were
equipped with functional long iron swords, bars and ingots began to circulate, partly as
a mean of exchange as well. Many investigated bloomeries attest the production of iron.
This was the situation especially in the Celtic parts of Europe (La Tène B/C periods, 4th
to 2nd centuries BC).
The Fully-fledged Iron Age involves the large scale production of iron which became
indispensable in the everyday life of the society. About 60 - 100 or even more kinds
of artefacts (armament incl. the protective armour, artisan’s tools and agricultural im-
plements, personal ornaments and gear, domestic utensils and masses of structural iron
- nails, clamps, dowels) illustrate the importance of that technical metal. Specialized
branches appeared among blacksmiths: swordsmiths, armourers, cutlers, toolmakers and
many others. Since the Late La Tène period (LC/D, culture of Celtic oppida) specific
outlines characterized the Roman, Medieval and post-Medieval civilizations.
THE EARLY IRON AGES 15
16 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 1: The sequence of Early Iron Ages in the Near East and Europe. Abbreviations: E
- early, IA - Iron Age, LH - Late Helladic, Submyc - Submycenaean, PG - Protogeometric,
G - Geometric, arch - archaic, class - classical, Cimm - Cimmerian, Scyth - Scythian, Etr
- Etruscan, H - Hallstatt, LT - La Tène, pre-R - pre-Roman, R - Roman.
THE EARLY IRON AGES 17
The transition from Bronze to Iron is an interesting chapter in human history. Both
metals and alloys were first used as prestige goods by members from prominent social
strata and were secondarily introduced as technical materials for the general population.

A remark has to be added. The monophyletic view presented here and submitted to a
criticism of sources was recently rolled over by theories concerning western and central
Africa (Aux origines de la métallurgie du fer en Afrique, 2000). Basing exclusively on a set
of radiocarbon data from several sites connected with iron objects, slags and occasionally
furnace remains, the chronology of the use and making of iron has been pushed back
to 1500 - 800 BC and declared to be an autchthon phenomenon. Nonetheless, the short
conclusion of the book quoted (Maes-Diop, o.c. 189 - 193) destroyed all critical approaches
to the sources, mixed facts and events together and resulted in a strict verdict: Africa, at
least its central part (Nigeria, Niger), developed iron making and working independently
and in a parallel way as in Eurasia and practiced it from the 3rd millennium BC (?). No
doubt that this should be the very subject of further analytical studies and discussions
among archaeometallurgists.
Chapter III

IRON AND STEEL

Iron (ferrum, Fe) is a silverish metal, atomic number 29 in the eight group of the
periodical system, atomic weight 55.85, Mohs hardness 3.5 - 4.3, melting point 1529 ◦ C,
ferromagnetic. It is malleable and weldable. In pure form it exists an an electrolytical
substance which is produced from an aqueous solution of ferrous salts or from iron pen-
tacarbonyl (FeCO5 ). It posseses a considerable affinity to hydrogen and beacause of that
it corrodes easily. The product is used in the iron powder metallurgy. It is evident that
such iron cannot be the subject of any consideration dealing with the early history of this
metal. This is reserved for technical iron containing small amounts of other elements.

Nomenclature

However, the modern normative nomenclature (early decades of the 20th century) calls
all the technical iron as steel, classified according to its carbon content. In addition, the
industry is familiar with alloy steel containing about 4% of other elements influencing its
properties.
Recently, the steel has is been produced by fining blast furnace cast iron. Different
conditions held sway in the sphere of the bloomery process yielding iron of pasty consis-
tence which was enormously heterogeneous in terms of its carbon content: the resulting
sponges or blooms varied in the content of this element from practically carbon-less metal
up to hypereutectiod parts (up to 1.7%C) with accidental occurence of cast iron droplets.
Nevertheless, the modern term ‘steel’ is even nowadays used inconsequently: up to re-
cent times an Iron and Steel Institute existed in London as did Stahl und Eisen periodical
in Germany, and the technical terminology involved the term wrought iron for malleable
mild steel. Curiously enough, the elevated carbon content did not hindered the use of
terms like cast iron, pig iron an the word ‘iron’ for various kinds of tools. Therefore,
the terminological game of technologists of the early 20th century seems to go beyond
its lexical sense. Both terms, ‘iron’ and ‘steel’ have their history which cannot be ommited.

Etymology of the words ‘iron’ and ‘steel’ in European languages

It is not without interest that, except for Greek and Latin, the germs for the word ‘iron’
denoted originally a non-ferrous material in all of the Indo-European languages. In the
Vedic passages ayas and Avestan ayanh mean copper or copper-based alloy which were
derived from aios in reconstructed primeval Indoeuropean. Hence it was transformed
into Latin aes and Gothic aiz (copper, bronze) and exists in the Old German adjecive
ehern. Iron, when recognized, was distinguished by attributes (syaman ayah in the Vedas,
kalayasa in the Sanskrit).
It follows that the ancient Indo-Europeans were secondarily acquianted with iron during
their ethnogenesis. The above mentioned Indoiranian stem appeared later in the West as
iron: Celtic ı́sarno, ancient Irish ı́arn, Anglosaxon ı́sern, iron, Old German ı́sern, German
Eisen.
IRON AND STEEL 19
A quite different root for the word ‘iron’ evolved in the eastern Indoeuropean Baltic
and Slavic territories. The stems gel-, dzel-, zhel- were also used originally for a non-
ferrous alloy (alluding to something yellow) and were in subsequent ages adopted for iron
(Russian, Slovak and Czech železo, Polish żelazo, Lithuanian gelžis, ancient Prussian gelso,
Lettish dzels. Thus, in the early centuries of the 1st millennium BC the Indo-European
peoples of the eastern, western and northern Europe renamed the technical metals in
question.
The South of the Continent, also inhabited by Indo-Europeans, Greeks and Italics,
witnessed more advanced situation in the field of metallurgy: the non-ferrous branch
used specific terms chalkos (a red metal), together with the above mentioned aes or
cuprum, copper produced in ancient Cyprus. The Greek sidéros and the Latin ferrum
were reserved for iron. The etymology of the Greek word is veiled but ferrum is of non-
European origin: it developed from the Etruscan ferzom which alludes to the Near Eastern
barzel, parzillu, iron. By the way, Hattians, Hurrians and Hittites pronounced the term
AN.BAR/parzillu as hapalki, hapalkinnu, a non-Indo-European word as E. Laroche and
A. Kammenhuber have explained. Later, the term ferrum developed into the Italian ferro,
French fer, Spanish hierro and Rumanian fier.

Almost simultaneously another variety of iron was recognized and was specified by name
- the harder (carbon) steel. In the Greek world it was the property of a sharp cutting-line
— stoma, stomóma, or the metal as such — adamas, chalyps, i.e. steel. In Latin the word
acies expressed the same notion as the Greek stomóma and and gave rise to the modern
French term acier, steel.
In the North, substantially later in the chronology of ironworking, another germ for
this material entered the terminology which is the base for our modern word. In the Old
English stele, stȳle, in Old High German stahal, later Stahl. This was in the course of
time transferred to Polish and Russian (stal, stal’). Only the Czech language has ocel for
steel.
In fact, the story of terminological development reflects, as well, the spread of iron
technology in the South and northwards.

Thus, iron and steel were distinguished in the ancient and recent past. In daily smith’s
practice the wrought iron (in modern parameters 0.02% to 0.3% C) was not hardenable by
rapid cooling whilst steel could be hardened by quenching and tempering (see Chapter V).

Physical properties of iron

This metal has been introduced above in the first paragraph as a chemical subject. In
that sense it does not occur as a natural substance. It occurs in the nature in three forms:
(1) As the iron ore used for smelting during the bloomery process and, later, in blast
furnaces characterizing the indirect process.
(2) As meteoritic iron with about 3% to 18% nickel, O.1% to 0.3 % phosphorus and
ca 0.4% cobalt in the kamacite variety and up to 60% Ni in the taenite variety. Some
preserved sideritic meteorites survived the fall and have been studied. They represent
20 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
fragments of destroyed interiors of celestial bodies (it is supposed that the nucleus of
the Earth, the nife, is of similar composition). Discussion has arisen about the use of
meteoritic iron in ancient cultures (the Near East, Egypt); the nickel content of certain
previously analysed samples is not considered as decisive (Piaskowski 1980; Photos 1989).
Nickel when present in small amounts in the ore of certain deposits transits completely
into the smelted iron and segregates in some spots containing up to 9% Ni. Nonetheless,
the ‘metal of heaven’ in ancient nomenclatures is especially worthy of attention in the
term the AN.BAR.GE6 , a ‘black metal of heaven’ occuring in Hittite cuneiform texts
in the same documents as AN.BAR i.e. iron. It is interpreted as meteoritic iron used
for the manufacture of weapons and ritual objects, sometimes in considerable quantities
(Siegelová 1984, 159 - 163). In general, the problem is far from being solved: the analyses
of ancient objects are not easily controllable any more and no nickel contents are given in
the texts involving the AN.BAR.GE6 . On the other hand, the Inuit Eskymos near Cape
York or on the island of Disko in Greenland stroke off fallen meteorites (Craddock 1992;
Buchwald 2001, 55 - 61, 2005, 13 - 38). Despite the arguments of some scholars the use of
meteorites for manufacturing weapons and amulets (which are not to be discussed here)
cannot be excluded, in particular when observations of meteorite falls are attested in the
documents of the ancient Near East (Bjorkman 1973).
(3) As terrestrial or telluric metal occuring sporadically in certain regions of Central
Europe and in the far North. Again, the Eskimos exploited terrestrial iron blocks found
near Ovifak in Greenland (e.g. Coghlan 1956, 177, after Allen).

Iron melts at 1529-1531 ◦ C and is malleable in hot and cold state, and, as mentioned
above, weldable. The principal modifications in terms of the arrangement of atoms in the
crystal lattice (Fig. 7) are body-centred α-iron (up to 906 ◦ C whichis magnetic up to 760

C and face-centred non-magnetic γ-iron (906-1403 ◦ C) the austenite of which absorbs
carbon (see the equilibrium diagramme, Fig. 8). As mentioned above, the carburization
of smelted bloomery iron is extremely heterogeneous: the product may consist in vari-
ous proportion in individual spots of the bloom as ferrite, ferrite-and-pearlite, pearlite
(eutectoid steel), pearlite-and-cementite (hypereutectoid steel, over 0.8% C) and even as
particles of cast iron (ledeburite, cementite, graphite, see Chapter XI).Ferrite is crystal-
lized α-iron with less than 0.02 to 0.03% C and comes into existence when the cooling
of the metal (less than 0.1%C) from the austenite below 906-721 ◦ is slow (Pl. I: 1).
In the grain boundaries of a metal with 0.03-0.04% C segregates the so-called tertiary
cementite (Fe3 C). Pearlite (pearle lustre in metallographic polished samples) is a struc-
ture consisting an eutectoid combination of ferrite and cementite lamellae (Pl. I: 2); it is
produced in various proportions during slow cooling of hypoeutectoid steel (0.05 - 0.8%
C). In conditions of annealing the lamellae coagulate to globular pearlite (see Chapter
11). In hypereutectoid steels (more than 0.8% and less than 1.7% C) the secondary ce-
mentite appears in form of white thin cells (network in pearlite grains when sectioned) or
small arrows penetrating into pearlite grains. Cementite (iron carbide) as such contains
6.67% C. When an overheated mild or medium steel is submitted to relatively intensive
cooling the ferrite precipitates as needles (in fact fission faces in austenite). The result is
a pearlitic structure with typical ferritic ‘arrowheads’. The texture is known under the
term Widmanstätten texture (Pl. I: 3). Such steel is noted for decreased strenght and
toughness and in modern technology is usually submitted to annealing to be recrystal-
IRON AND STEEL 21

Figure 2: Atomic arrangement in the cubic lattice in iron: a - b body centred α-iron, c -
d face centred γ-iron. After Scott (with references).

lized. In archaeological steel objects the Widmannstätten structure appears in certain


parts of some artefacts (Pl. I: 3). (The iron and nickel alloy of sideritic meteorites dis-
plays usually the Widmanstätten structure as well). Should the steel be cooled rapidly,
other structures appear on metallographically examined samples. The above mentioned
and instable features as well are briefly discussed in Chapter XI.

Recent classifications of iron and steel

The material of the ancient and traditional smith was carbon steel containing sometimes
smaller amounts of phosphorus delivered from certain ore deposits (see e.g. Piaskowski
1965a and 1973; Nosek 1991) similarly as arsenic (Piaskowski 1984); both make the metal
slightly harder and more brittle and hinder the absorbing of carbon in steel. They often
segregate in welding seams. In modern practice carbon steels are classified according to
the carbon content.
The softest steel coincides with the material called wrought iron containing less than
0.03% C. It particularly found application in mountings and fittings and, above all, in
construction of tool bodies and backs. In modern times wrought iron has been known as
metal bands and as the structural steel with various cross-sections (I, U, T, L etc.). A
phosphorus content of up to 0.1% P is tolerable (otherwise the content of this element
in modern steel should not exceed 0.04% P). The classification of steel according to its
carbon content is not strictly unified. These containing less than 0.2% C (involving
the above mentioned kinds of wrought iron) could be considered as mild steel, whereas
medium steels range from a content of 0.3-0.5%C, hard steels (0.6% up to the eutectoid
material with ca 0.8% C) and very hard hypereutectoid steels (more than 0.9% and less
than 1.7%C). Another classification recognizes mild steel (up to 0.15% C), medium steel
22 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 3: Equilibrium phase diagram iron - carbide. Left: iron - cementite system; right:
eutectoid region in the iron - cementite diagram. Scott after Samuels.

(0.15 to 0.3% C) and hard steel (more than 0.3%C). All of the kinds may be encountered
among archaeological objects.
A different family is that of alloy steels which contain other specific intentionally added
elements (also in combinations) which give name to the material: chromium, chromium-
molybdenum, chromium-silicon, chromium-tungsten, -vanadium, -nickel, cobalt-chromium,
silicon, manganese, silicon-manganese, molybdenum, tungsten (wolfram) etc. These con-
stituents and their amount influence the properties of the material according to pro-
posed purpose and cause different behaviour of crystalline phases when heat-treated. The
bloomery period was ignorant of any intentional alloyed steel.
Other classifications concern different aspects, e.g. desoxidation degree (killed and rim-
ming steel), purity (plain, high-grade), purpose of use (structural, reinforcing, magnetic,
tool, spring, free cutting and high-speed cutting, stainless etc. steels). Or according to the
shape of semi-products: strip, hoop, sectional (I-steel or joist, U-iron, T-iron, see above).
These kinds were, of course, unknown in prehistory, antiquity and Middle Ages. Instead,
there were various types of bars, rods, blanks and billets, square- and round sectioned or
in the shape of stylized artefacts. This theme is discussed in the next Chapter IV.
Chapter IV

THE BLACKSMITH’S STARTING STOCK

The smith who intended to produce artefacts and other usable iron things had to reach
for material which could be worked, either from primary imported blooms (see Pleiner
2000, 230 - 250) which he had to divide and adapt for further processing, or from suitable
pieces of iron or steel already formed from bloom or iron sponge: typologically inexpressive
blocks, bars, rods (Fig. 9).
A number of such pieces have been found among finds complexes that have been un-
covered in layers of settlement centres of all periods. Early examples should be brought
to attention from the Celtic oppidum of Manching in Bavaria (Jacobi 1974, 253 - 254,
pl. 77: 1502 to 1504, 1508, pl. 78: 1537 to 1542). Straight or bent square-sectioned
bars as well as small iron pieces and fragments came to light among early Roman finds
from Magdalensberg in Austria (Dolenz 1998, 232 - 236, pl. 89 amd 90). Some of them
may represent certain kind of semi-products or scrap suitable for recycling. The bent
square-sectioned rod B6 (1.32kg, Fig. 2.1: 14) was heterogeneously carburized (0 - 0.8%
C, see Straube 1996, 131 - 134, figs 44 to 46).
However, within European archaeology a number of kinds of virtual shaped iron ingots
and bars have appeared since the early periods of the Iron Age.

Bipyramidal ingots

The term involves iron bars whose development marked a part of the European economy
in course of more than half a millennium (late Hallstatt and at least the entire La Tène
periods). A terminological note should be added: The bulk of these ingots is spread over
Celtic territories which mainly extend troughout SW Germany where the term Doppel-
spitzbarren or Spitzbarren is used. English refers to them as double pointed bars (stressing
their tips) or as spindle-shaped bars (but no spindle is edged). The earliest models are
not double pointed but the bodies of all variants are bipyramidal so that this term corre-
sponds with their shape.

Early models of bipyramidal bars

In 1860’s a stone-walled treasury was uncovered at Khorsabad (ancient Dur Sharukin)


in Iraq (Place 1867/70). It was 5m long and 2.6m wide and up to 1.4m filled with iron up
to a depth of 1.4m: about 160 tonnes of bipyramidal ingots, rings, chains and other imple-
ments. Apparently this was a part of spoils and taxes in iron collected by Neo-Assyrian
rulers from Anatolia, Urartu and Syria in course of the 9th an above all 8th century BC
(Pleiner and Bjorkman 1974, 291, figs 6 and 7). Most of these bipyramidal bars were lost
during the wrecking of a sailing ship which was transportingd them to France. Several
pieces survived being kept in Louvre (Fig. 10: 1 to 10) and some others are said to
have been deposited in the Iraq Museum at Baghdad. Later, American archaeologists
uncovered additional bars (Oriental Institute Museum, Chicago). Apart from Khorsabad
several bars of that type were found in Nimrud (now in the British Museum) and at Susa
(Louvre), see Fig. 10:2. 11 to 13, (Curtis et al. (1979, 376) add a note on a similar object
24 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 4: Various iron stock. 1 - 11 Manching, Bavaria, La Tène period oppidum; 12 - 15


Magdalensberg, Austria, early Roman. After Jacobi and Dolenz.

from Karmir Blur. Fluzin (2002, 71 - 72, pl. III: 23) shows a lighter bipointed ingot from
Carthage (N Africa, 4th - 3th centuries BC); its round hole is in the centre of the body
(lengtht over all 20cm, weight 1.77kg).
As stated above, the ingot bodies are of bipyramidal shape but not bipointed. The
form is asymetrical - one tip resembles a bent bird’s beak, the other is flat like a fish-tail.
Near the beak a round hole is drifted into the metal (Figs. 10 and 11: 1 - 3); possibly it
served for easier transport having been passed through with a rope. However, specimens
without any hole occur, especially in European regions. The length of the Assyrian ingots
varies between 30cm and 50cm (the centre being some 7cm x 14 cm thick). The weight
was 4kg to 12kg (more than 20 minae) but the heaviest piece preserved in the Louvre
does not exceed 9kg (Pleiner and Bjorkman. l.c.). Two bars from Khorsabad and the two
from Nimrud were metallographically investigated (Pleiner 1979b, Curtis et al. 1979).
The metal was relatively pure but very heterogeneously carburized with spots of ferrite
and pearlite up to eutectoid values (0.8% C). The forging must have been finished at
lower temperature (about 710 ◦ C) and decarburization of thinner parts must be taken
THE BLACKSMITH’S STARTING STOCK 25

Figure 5: Neo-Assyrian bipyramidal iron bars. 1 - 10 Khorsabad, 11 Susa, 12 - 13 Nimrud.


1 - 8 courtesy of the Louvre: 1 AO 24106, 2 AO 24107, 3 AO 24111, 4 AO 24110, 5 AO
24112, 6 AO 24108, 7 AO 24109, 8 AO 24113. 9 - 10 courtesy Oriental Institute Museum,
Chicago, 9 A 12261, 10 A 12462. 11 Courtesy Louvre SB 9159. 12 - 13 courtesy of the
British Museum, 12 N-1963, 13 N-962 (chopped piece). After Pleiner and Bjorkman.
26 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
into account. The above mentioned specimen from Carthage (Fluzin 2002, l.c. ) was
cut by a complete longitudinal section which shows cracks at both tips; no structures are
described.

In Europe, bipyramidal bars of beak-and-tail type appeared as well. They date from the
late Hallstatt period (6th - 5th centuries BC). The geographical distribution covers the
central parts of the Continent leaving a wide gap as to the finds in the near East, although
some of them seem to have been formed after the Neo-Assyrian model; they are a little
smaller and lighter. A neglected find is that of 25 bipyramidal ingots with holes (25cm
to 37cm long (Fig. 11: 2) deposited as votive gifts in front E of the temple and close to
the temenos of Neoptolemus at Delphi (Pleiner 1969, 17, and fig. 5: 7, after Perdrizet).
Similar bars are said to be found at the early Greek colony of Naukratis in the Nile delta
(Perdrizet 1908, 213).
A Danube river find of Dunapentele-Dunaújváros (S of Budapest) deserves the atten-
tion in this respect comprising 6 beak-and-fishtail bars with holes (length 33cm to 51cm,
weight 1.12 to 1.76kg, Fig. 11: 1, see Szabó 1966). Two bars with holes and beaks con-
tained a Hallstatt period hoard from Biskupin, a palisade settlement in W Poland (Fig.11:
3, see Kostrzewski 1953, figs 46 and 47). About of the same date are the halves from
the Býčı́ skála deposit (Moravia) and from the hoard found at Leipzig-Wahren in central
Germany (weight about 3kg). Both were evidently hot-chiseled which means that they
were used as starting material (as the Nimrud case was) but nothing is known about their
complete shape (Fig. 7 and 8, see Pleiner 1958, 81 - 82, Fig. 16; Peschel 1980). For exam-
ple, another bipyramidal ingot from Maszkowice (S Poland), 28cm long and 0.76kg heavy,
is noted for its long and slim point but it has no hole (Fig. 11: 4, see Cabalska 1964). The
metallography of that piece yielded information about the heterogeneous distribution of
its carbon (up to 0.9% C) and slightly elevated P-content (0.085%, see Piaskowski 1977).
Similar in shape but with slightly curved tips are 5 unstratified bars from Witów, Poland
(individual pieces about 0.8kg, Fig. 11: 5) are similar in shape but with slightly curved
tips; they were unevenly carburized from 0.05% to 0.4% C, one of them (No 5) up to 0.8%
C; the phosphorus content was rather low (0.041% P). The examinations were carried out
by Krupkowski and Reyman (1953, 58 - 59, fig. 9) and later by Piaskowski (1962).
Beak-pointed (órnéón týpos after Kleemann) are known from the proto-Celtic milieu
in western Germany as well. A typical complex of that kind is a Hallstatt period hoard of
23 bipointed ingots from Aubstadt south of Mainz (Fig. 11: 9, see Kleemann 1966). Its
total weight was nearly 100kg, the individual pieces were of some 3 to 4kg of weight and
were 40cm to 60cm long. Another example of a similar type was found in a burnt-down
house in the centre of Heuneburg near Hundersingen; although it corresponds with those
from the Hallstatt type (Fig. 11: 6), the hut in question dates from the early La Tène
period (Kimmig and Gersbach 1971, 54). The ingot was 42cm long, weighed 7. 5kg and
was metallographically investigated by C. Tölg (ibidem, o.c. who carried out complete
transversal and longitudinal sections). Both points and surfaces were steely (0.2 - 0.6 %
C) whilst the centre contained 0.04 - 0.06% C. The heterogenity was explained as a result
of welding together of smaller pieces. More robust bipyramidal ingots were published from
Armsheim near Mainz said to be, on the basis of discovered sherds of Hallstatt period date
(Weiershausen 1939, 198, fig. 63). A river find from the Sâone near Mâcon comprised
THE BLACKSMITH’S STARTING STOCK 27

Figure 6: Early bipyramidal ingots in Europe. 1 Dunaújváros, Hungary, 51 cm long,


around 500 BC; 2 Delphi, Greece, 25cm; Late Hallstatt period: 3 Biskupiń, Poland, two
ingots representing a hoard, 21cm; 4 Maszkowice, Poland, 28cm; 5 Witów, Poland, ingot 2
from a hoard; 6 Heuneburg, SW Germany, 25cm; 7 - 8 halves: 7 Leipzig-Wahren, central
Germany, 26cm ; 8 Býčı́ skála cave, Moravia, from a deposit, 10cm; 9 Aubstadt, W
Germany, early La Tène period, hoard (reconstructed). 1 after Szabó, 2 after Perdrizet,
3 after Kostrzewski, 4 after Cabalska, 5 after Krupkowski and Reyman, 6 after Kimmig
and Gersbach, 7 after Peschel, 8 after Pleiner, 9 after Kleemann.
28 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
some bipyramidal bars with slim and bent points another one was dredged from the Oise
river (Fluzin 2002, 71 - 72, pl. III: 22 and 25)
A different variant ascribed to the late Hallstatt/early La Tène periods and to the
Celtic people who inhabited the east of France comes from Mont-Lassois, Côte d’Or,
and Strasbourg (7 items): short and squat with straight horns (France-Lanord 1963)
and signalizes the coming of a new member in the family of numerous La Tène period
bipyramidal ingots. The Mont-Lassois piece was 28cm long, one of those from Strasbourg
25.8cm: the weight of the latter was 6.45kg. According to research by France-Lanord
both were heterogeneously carburized - the Strasbourg example contained 0.01 to 0.62%
C, O.12% Mn and 0.495% P.

Celtic bipyramidal ingots of the La Tène period

In earlier periods the custom of shaping iron into forms of sporadically occuring bipyrami-
dal ingots incorporated, in the earliwer period, various lands in central Europe including
those which cannot be included to the pre-Celtic domaine (Hungary, Poland, central Ger-
many). Nonetheless, gradually the lands which already by Hecathaeus were denoted as
Keltiké and by Herodotus (IV. 49) as inhabited by Celts, became the scene were the bulk
of finds could be registered.
During the period of about 500 to 50 BC these ingots presented themselves as a flood
in certain territories of the European Celtic domaine roughly defined by the Mosel, Lahn,
Altmühl, Isar, upper Rhône rivers in the lands of the Helvetii, Sequani, Vindelici, Raeti
- in eastern France, Württenberg, southern Bavaria. They occur sporadically in Gaul,
Normandy, Brittany (even once in southern England), Hesse, and the Alpine regions.
More than 700 pieces were recorded, mostly in several tens of hoards. Most of them are
undated but despite the original opinion by Kleemann (1961) the chronological framework
covers, in the light of several dated cases, the entire La Tène period: some individual pieces
from the later Roman times have to be regarded as accidental finds.
Four varieties come into question. The first stems from the earlier model with long
curved beak-shaped points, the second is more or less symmetrical and is noted for straight
long horns; the third is short and squat like the Mont-Lassois and Strasbourg items; the
‘fish-tailed’ piece from Manching is an exception to date. The first three variants were
present in the second hoard from Sauggart which proves that the hoard might have been
composed of ingots produced in different smithies. The locality of Sauggart, Württenberg,
is one of the most important (Fig. 12). The find from 1875 contained 15 bars, the second
hoard was found in 1934 involving 24 items, 155kg of total weight and consisting, as
mentioned, of three sub-types (Paret 1934, 62; Fry 1953, 50). The length of individual
pieces oscillates between 30 - 86 cm (the one with extraordinary long horns even measured
140cm. One of the bars was chemically and metallographically examined by Fry (o.c.).
The results of investigations carried out on different ingots are briefly mentioned below.
The hoard of Bechtheim (northwards of Worms, W Germany) consisted of 60 pieces of
the long varety, that of Kaisheim (upper Danube valley) of 26, and 17 bars were found
in 1961 (dredged out but possibly heaped together) at Renningen west of Stuttgart (1.5 -
4.6 kg; one item was metallographically investigated: Zwicker 1967, see also Fundberichte
aus Schwaben 18/II 1967, 75). Nine pieces representing 42kg of iron had a hoard from
Rodalben (eastwards from Saalburg, W Germany).
In Switzerland a certain cummulation of bipyramidal ingots can be observed in the
THE BLACKSMITH’S STARTING STOCK 29

Figure 7: Celtic bipyramidal ingots. Sauggart, Germany, selected varieties of ingots from
the second hoard. Lengths: 1 53cm, 2 45cm, 3 47cm, 4 120cm, 5 - 6 ca 30cm. After Allen.
30 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
environs of the lakes Neuchâtel and Bienne (Biel). At Schwadenau 16 ingots came to
light (Serneels 1993, 168 - 169, fig. 189, after Müller).
Two examples from the northern Celtic periphery came from the oppidum of Steinsburg-
Kleiner Gleichberg in central Germany (one was examined), and another two were found
in a house within a Roman military station at Homburg v. d. Höhe north of Frankfurt/M
(one is gone lost, the other weighs 5.52kg, see Kolling and Schähle 1968); the question
arises about the origin of the ingots (old stray finds deposited later?). According to Fluzin
(2002, 71 - 72, pl. III: 22 and 25) a bipyramidal ingot from Coulmier-le-Sec in France (Côte
d’Or) could also be of Gallo-Roman date (16.3cm long, 4.75kg); no details are reproduced.
There are also three bipyramidal bars with bent horns from Saint-Jean-Trolime, Finistère,
Brittany (Vercingetorix et Alésia 1994, 53 - 55). (Note: A La Tène period fortification
at Plavecké Podhradie-Pohanská in W Slovakia yielded six iron bars which have pointed
tips but no bipyramidal bodies; they rather resemble rods. No detailed information is
available after the preliminary report by Paulı́k in 1970).
Metallographic analyses of bars from Renningen, Kaisheim, Ay, Steinburg and Touf-
freville (Zwicker 1967; Rädeker and Naumann 1961; Hanemann 1930; Fluzin o.c., pl.
III: 24 and 27) show not only that the metal was heterogeneuosly carburized during the
bloomery process (0.05% up to 0.5%, and even up to 1.3% C) but also that the bars were
each welded together from several blooms - some of the welding seams even appear on
the ingot surface, others on polished sections where considerable segregation of phosporus
may be observed (Renningen up to 0,7%). The joining of several blooms or blocks to
bipyramidal ingots seems to be logical since the La Tène period bloomey furnaces could
hardly deliver blooms heavier than 2kg - 3kg and during the reheating and forging at
least 1kg or more was lost again by reoxidation (most of these ingots weigh 4kg - 7kg).
A question remains: why must have been several blooms joined together to be later la-
boriously divided by hot chiselling when a part of the metal was intende to be used for
other purposes? The forging of points, which were often very thin, may be explained as
a visible demonstration of excellent malleability.
And why were tonnes of valuable iron taken out of circulation and use is a question
which has little chance of being solved. Ritual reasons cannot be excluded in a world so
heavily tainted with cults and superstitions. Possibly, these moments might have lead to
economical collaps of a technically prospering society.
Other quantities of iron have been deposited in hoards and never used again - iron in
form of elongated bars. What could not be used again were goods that had disappeared
during transportation : bipyramidal or ‘modèle au double pointe’ and massive flag ingots
as well were carried by ship which sunk at Capo Bellavista at the eastern coast of Sardinia;
the date is given between the late 1st century BC and 1st centuy AD (Feugère and Serneels
1998, 253, fig. 3. 256, after Parker).
When closing this section devoted to bipyramidal bars (Fig. 13: A) it is absolutely
necessary to deal with the finds from layers of the Manching oppidum in Bavaria (Fig. 14).
First, two kinds of bipyramidal ingots were uncovered: hot-cut halves of regular ingots
(with traces of tongs graspings), then a bar called ‘Manching type’ which resembles fish-
tail ingots, 44cm long; a rectangular ashlar-shaped iron block (13.7cm x 7.3cm) which
had apparently (Fig. 14: 1) been through the working processes - again, with deep tong
grip grasping traces (Jacobi 1974, 337, pl. 76: 1498, 1499, 1500, 1501). And what has
already been stressed: within the same site examples of elongated flat bars which are
sword-shaped (ibidem, pl. 77: 1505 - 44.6cm and 1506 in two fragments, (Fig. 14: 6 and
THE BLACKSMITH’S STARTING STOCK 31
32 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
7). The mutual presence of different forms at one La Tène period centre is represented
by the fact that another category of iron billets called saumons d’épée, schwertförmige
Barren, taleae ferreae denoted as ‘currency bars’ were in circulation.

Figure 14: Manching, Bavaria. Iron ingots and sword bars (6 and 7) from the Celtic
fortified centre. After Jacobi.

Taleae ferreae

The Latin term denotes pegs or bars made of iron and paraphrases that used in the
Commentarii de bello Gallico by Caesar (5.12) which will be discussed below. The group
involves tanged sword-like pieces as well as socketed bars (some of which were interpreted
as ploughshares).

Tanged sword-shaped bars

These bars are approximately 40cm to 50cm long and about 3.5cm wide weighing some
0.75kg (Fig. 15: 1). Their tang is shouldered and pointed taking about one third of
the total length and its root bears traces of an original rolling. The ‘saumons d’épée or
Schwertbarren resemble, in fact, starting bars for making sword blades. Whether it was so
remains unproved but it was possible: sword blades have been experimentally forged from
THE BLACKSMITH’S STARTING STOCK 33
replicas of such bars of the La Tène type sword blades were experimentaly forged (Pleiner
1993, 71 - 77). The distribution spread over Switzerland and eastern France. Apart of
an example from Port-Nidau which served as a model for the above mentioned replicas a
hoard hs been announced from Bern-Tiefenau (Müller 1990). Some objects were rescued
from the river Sâone near St. Marcel as well as the 12 pieces dredged out from the river
Seurre, Côte d’Or (see Vercingetorix et Alésia 1994, 53 - 55; Kruta and Szabó 1979, fig.
77). A piece 34.5cm long was found at the Celtic Heidetränk oppidum in Taunus, W
Germany (Müller-Karpe 1977, fig. 8: 17). In Appleford, Berkshire, England, two bars
were rescued in 1967 from a bundle of six pieces one of which is very similar to the sword
bars so that Brown (1971, fig. 2: right) dallied with the idea that it was virtually a sword
fragment that had been added to the other socketed bars.

Socketed bars of the Wérimont type

Better preserved bars reach 30cm - 50cm in length and their blade may be tapered to a
point. The head part is rolled to a more or less closed socket (Fig. 15: 2). The name
originates in bars found in the Bois de Wérimont near Namur in Belgium (Schäfer 1984,
after Mariën). Four examples were found at La Tène, Switzerland, and the type also is
known as Wérimont-La Tène. It spreads over northern France, Belgium, northern Ger-
many up to the Saale river, then over eastern France, Switzerland, and Bavaria up to the
Altmühl river (about 30 sites in total).
A hoard of 9 bars was uncovered at Säffig near Koblenz, W. Germany, on the territory
of a La Tène period settlement (LC2/LD) destroyed in 1980 - 1982 (Schäfer 1984). They
are damaged, 15cm to 20cm long and 1.2cm to 1.6cm wide and belong to the smallest
and lightest of their kind (ca 0.05kg). The fragment of bar 1506 from the Manching
oppidum (Jacobi 1974, pl. 77) may be counted amongst the type. Further north museum
collection research brought the trace of a La Tène period hoard from Ochtrup (1891)
which comprised 4 bars and smithing and other tools (Wilhelmi 1977; Polenz 1980); six
sword bars (25cm - 46cm) are known from a hoard nearby Münster-Stadkern (ibidem).
Outside the territories of the spread the following recent finds deserve attention: the
eight socketed long bars (70.5cm to 80cm, 1.165 to 5.145kg, Fig. 2.7: 4)) from Montans,
Tarne, S France, 2nd - lst centuries BC (Martin and Ruffat 1998); another find has been
announced in the neighbouring Rabestens (o. c. and Fluzin (2002, 71 - 72, pl. III: 21)
depicts an identical socketed bar from Aulnat, Auvergne, S France (0.737kg, late 3rd
century BC). These bars deviate by length and weight a little from the Wérimont type
and seem to be earlier than the British ‘currency bars’ (see below).

‘ Schwurschwerter’

The so-called ‘swords of oath’ represent another group of long bars. The origin of this
nickname inhers in a tale connected with a find in the courtyard of the Wartburg castle
in central Germany in 1845/46 (Fig. 15: 3). According to Ritzgen a bundle of 13 long
bars was deposited in a rock cavity and joined together by a wire (what sort and origin?).
A tale arose that these bars, 67cm to 97cm long (0.235 - 0.75kg) were supposed to have
been objects of oath or swearing but Götze (1928) already classes them as Iron Age bars.
In spite of differing shape he identified them with the ‘currency bars’ of England and
did not exclude that they even been imported. The blunt side edges of the Wartburg
34 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
bars are slightly tapering. At a point 1/6 along the total length which varies between
84cm and 97cm (two examples are shorter and broken) is an uneven contraction dividing
the ‘blade’ an the pointed ‘hilt’ part (Fig. 15 :3). Two items which show a better state
of preservation are suspected to be additional replicas of the 11 original bars (see also
Weiershausen 1939, 204 - 207, fig. 68); there is a note concerning the chemical analysis
of rust by Hanemann which seems to be irrelevant due to contamination with the soil
components.
Similar bars were found at Heiligenstadt (4 examples), Wichdorf (6 pieces), Gettenau
(10 pieces), Münster, all of them in the northern periphery of Central Germany (Weier-
shausen o. c., 207 - 209, figs 69 and 70). Later, the same author (idem 1942) summarized
the ‘Schwurschwerter’ finds and expressed his verdict: they are ploughs and coulters,
which is supported by the construction of the plough from Dabergotz (o. c., 89, fig. 8)
for which no detailed information has been published. The possibility of using or stylizing
the long bars as ploughs has been expressed in discussions on the Iron Age bars in Britain
which are the subject of the next section.

‘Currency bars’ in Britain

Talking about the story of elongated bars of the taleae type we cannot avoid quoting
the phrase in Caesar which uses this word and turns the head of scholars. The relevant
place in Caesar’s De bello Gallico (5.12) reads: Utuntur aut aere, aut nummo aureo, aut
taleis ferreis ad certum pondus examinatis pro nummo. The Gauls in Britain ‘use either
copper or coins of gold or peg-like iron billets of fixed weight as money’.
It is clear that this iron played a role in trade. Whether it was used, in addition,
as currency is a matter of endless discussions. A piquant fact arises that in southern
England which was visited by Roman troops in 55 and 56 BC enormous numbers (more
than 1500 pieces) of elongated socketed bars were found which were identified in the early
20th century with Caesar’s taleae and named ‘currency bars’ (Smith 1905). Reginald
Smith tried to find a weight system among the above funds and proposed 0.309kg as
the basic unit, and then halves (0.155kg), doubles (0.618kg) and quadruples (1.236kg).
The ‘currency bars’ of Britain were submitted to further research (classifications by Allen
1967; Hingley 1991; Crew 1994). Critical views were expressed about Smith’s treatise.
Objections appeared above all against the weight system taking into account that the
dimensions could have been more easily calculated than the weights. The bars were
mostly explained as blanks for making sword blades (Hulme 1933, 67; Tylecote 1962, 211;
Brown 1967). British authors gradually inclined not to interpret the bars as currency
units in the sense of Caesar’s phrase arguing mostly that it was impossible for smiths to
produce bars of four distinct weights. Crew (1994, 346) in his detailed study came to
conviction ‘that the old concept that the bars had a series of standard weight and that
they were used as a currency in the conventional sense of that word, is clearly no longer
tenable’.
The taleae ferreae of southern Britain were marginally discussed by archaeologists
and archaeometallurgists on the continent in connection with similar semi-artefacts in
various regions of Europe (Götze 1928, treating the ‘Schwurschwerter’ from Wartburg;
Weiershausen 1939, 203 - 204, rejected their function as currency and explicitly (idem
1942, 85) defined them as plough-shares). Attention has also been paid to the British
bars in other works (Pleiner 1962, 68 and 1980, 251; Jacobi 1974, 253; Doswald 1994b,
THE BLACKSMITH’S STARTING STOCK 35
334).
The following lines will be devoted to an abridged passage dealing with ancient British
‘currency bars’ in the light of several aspects. First, the circumstances of finding should
be briefly touched on. A great deal (90% according to Crew) was uncovered in the 19th
century. In the majority these were deposited as hoards: Meon Hill 1824 - 394 items in
the centre of the fortified site; Malvern 1856 - 150 objects sintered together and a second
hoard of another 150 bars; Bourton-on-the-Water 147 pieces in remains of a box; Ham Hill
1845; Holme Chase 1870 - about a dozen bars hidden on and under a stone; Maidenhead
1894 - a bundle of 7 - 8 objects. These important complexes suffer by two facts: an
uneasy control of the original situations and of reports and on the state of preservation
(dimensions, weight, completeness of individual items). Crew (1994, 345) stressed the fact
that the greater part was found in hillforts and that the bars in hoards were of the same
variety, whilst they could be of different subtypes in settlement layers (Danebury). The
reason of hiding masses of iron escapes as it does in cases of hoards in general - hiding of
values, sacrifices?
The other aspect is the shape. The Latin word talea denotes a wooden peg so that
the attribute ferrea evidently indicates an elongated object of iron - a bar. This is in
full conformity with archaeological features. Formerly, roughly three main varieties were
distinguished among the British ‘currency bars’ were roughly distinguished three main
varieties: sword-like bars are slightly tapered (Fig. 15: 6), spit-like object are narrower
(Fig. 15:7) with somewhat parallel running sides and those which resemble, to certain
extent, the plough-shares (Fig. 15: 8). Crew (1994, 346 - 347, 349, fig. 1) classified
these bars into five varieties, each of them differentiated by the form of mostly overlapped
sockets (some rolled around a wooden rod (the traces of which have been preserved in
several cases, among Beckford and Meon subtypes) or rudimental and short. The sockets
should demostrate the good malleability of iron showing no cracks. The tips, when not
cut out, are oval and some of them are folded back and welded (Llyn Cerrig Bach and
Orton, see Crew and Salter 1993, 26, fig. 1). The individual types are believed to be
products of individual workshops or workshop complexes. A little larger in form are the
bars classed as the ‘plough-share’ type. The sockets are usually open with rudimentary
laps. The so-called hook billets (Fig. 15: 9), massive objects of the Meare type are
relatively rare and deviate from the elongated bars; they are trapezoidal, short and bent
into a narrower tip and they are heavy - 1.2 - 1.64kg (Crew, o. c. fig. 2: U; Fell 2003).
An important aspect is that of the correlation of the bar dimensions with their weight.
They are affected to a great extent by the fragmentariness and corrosion of the examples.
Longer and shorter pieces occur in all varieties and sub-types. The longest example is
represented by the find from Hod Hill - 87.6cm whilst that of Littleton belongs amongst
the very short examples (32.3cm) but it is the heaviest (Tylecote 1986, 146). The width
of bars varies between 2.5cm - 5.5cm and the thickness in the preserved state between
3mm - 5mm but the Maidenhead and Datchet types are 6mm - 8mm thick. There is no
correlation either in the dimensions or in the weight. The values are jumping. It seems
that the smiths producing ‘currency bars’ had not any balances at their disposal as the
Romans did; they rather used their eye to propose the rough dimensions of their bars.
This speaks against the ‘examined pondus’. Nonetheless, certain weight categories must
have existed the heaviest bars being observed in the Maidenhead, Coffinswall, Park Farm
and Datchet groups (0.8 - 1.451kg); Orton, Glastonbury, Llanstephan and Datchet types
include much lighter finds (0.3 - 0.5kg); explicitly lighter are those from Meon - 0.14 -
36 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 15: La Tène period elongated iron blanks and ‘currency bars’. Sword-shaped:
1 Port Nidau, Switzerland, 40cm, Wérimont type: 2 Ochtrup, Germany, 10cm.
Schwurschwerter: 3 Wartburg, Germany, 75cm; 4 Montans, S France, 90cm; British
‘currency bars’: 5 Salmonsbury hoard, England, 76cm; 6 sword-shaped bars with varying
socket form; 7 Spit-shaped bars; 8 plougshare-shaped bars. Hooked ingot: 9 Houghton
Down, 32cm. 1 after Pleiner, 2 after Wilhelmi, 3 after Weiershausen, 4 after Martin and
Ruffat, 5 - 8 after Crew, 9 after Fell.
THE BLACKSMITH’S STARTING STOCK 37
0.16kg. The Ely, Beckford groups can be regarded as medium (some examples. from
Danebury weighed 0.4 - 0.7lkg approximately). In the light of these data Caesar’s or
his news gatherers’ information about examined iron currency ‘ad certum pondus’ is not
cofirmed among the taleae. An idea of Götze’s (1928), namely that the weight of the bars
could have been examined per bundles remains an interesting speculation.
Chemical analyses have so far realised yielded highly variable but rather high carbon
contents within the Gretton, Beckford and Coffinswall group types whilst the Danebury
and Datchet pieces are of soft steel. The main criterion indicating the exploitation of
certain ore resources is their low or elevated phosphorus content Low phosporus bars
include the Beckford, Meon, Datchet and Llyn Cerrig Bach types which may originate
in the Forest of Dean area; Danebury, Gretton and Orton groups revealed phosphoric
metal indicating the use of Northamptonshire siderites and other local phosporus-rich
deposits. Crew (1994, 346) sees a correlation between the shape of the bars and its
metal. Some bars of the Gretton group show variable or high arsenic contents. From
some sites also come with a high Co and Ni content. The carburization of bars submitted
to metallographic investigations was high (possibly surface decarburized during heating
operations) or uneven (ferritic zone altering with pearlite - a picture which is, by the way,
typical for a great deal of bars of any kind). The Bourton-on-the-Water example (0.2
to 0.8% C) was a faggot of heterogeneously carburized bands (Tylecote 1986, 210, a full
cross-section see pl. XXI).
The place in Commentarii de bello Gallico (5.12) concerning the taleae ferreae, though
so much discussed and charged with various doubts, finds a conspicuous counterpart with
the bar finds. They represent material blanks stylized to specific customary shapes. They
circulated (and were hidden as well) as a form of trade iron purchased against values
related to its quantity (may be not explicitly to its weight).
At the time of Caesar’s raid to Britain in the 1st century BC the mode of using iron
currency in shapes of stylized tools was not new. Ancient Greece was already familiar
with this kind of means of exchange in the early first millennium BC.

Ancient Greek barter iron


The situation of the late Geometric, Archaic and Classical Greece (8th to 4th centuries
BC) has to be mentioned in terms of the circulation of iron. The increase of iron pro-
duction enabled to the development of different kinds of trade, barter and currency iron
which cannot be held for mere ingots.
The ocurrence of iron currencies in the form of stylized implements is very interesting of
this. Gradually, certain types became established in Greece (Pleiner 1969a, 15 - 17, with
references). Tripods (tripoda) and presumably anchors (ankyrai) and sickles (drépana,
Fig. 16: 2) were used as a means of exchange. Tripods as currency are mentioned in the
Gortyna Laws (I.C.IV.72). Anchors could have been a kind of harbour tax. About 50
sickles, evidently not normal tools, were discovered in a grave of the 5th century BC in
the island of Rheneia, opposite to Delos, and other items at Perachora.
However, the most important were the iron spits called obeloi or obeliskoi which gave
the name to modern Greek currency (obolos, see Gansiniec 1956). Originally they were
presumably used for roasting animal offerings in temples. The obeliskoi were used espe-
cially in the Peloponnese but they also appeared in other regions. Six of these iron rods
could be grasped in the palm of one hand (drachmé). This currency was evidently in
38 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
circulation in local markets in Greece before struck silver coins were introduced, and they
survived simultaneously with them for a time.

Figure 16: Iron currency in ancient Greece. 1 struck coin from Argos, 4th century BC;
2 stylized sickles from Perachora; 3 - 4 bundle of iron spits and the huge beam from
Argos-Heraeum, 1.12m; 5 iron spit from a warrior grave at Argos. After Pleiner.

According to tradition, Phaedon, one of the tyrants of Argos (mid-7th century BC)
withdrew the spit currency from circulation and dedicated it to Hera and replaced it with
silver coins. A large bundle of iron spit was discovered during the American excavations
at Argos in the late 19th century (Fig. 16: 3). They were tetragonal in section and
117cm to 130cm in length. Together with this bundle a huge specimen 120cm long and
8.8cm square in section (Fig. 16: 4), weighing 73kg, which was equivalent to 180 normal
obeliskoi. Assuming that the cost ratio to 1 : 400, the bundle of spits or the huge obelos
would have been one silver mina introduced by Phaedon (i.e. 30 drachms).
It would be pleasant to believe that the splendid Argos find was the famous offering of
Phaedon which was a symbol of his currency reform. However, there are many parallels to
this votive deposit of iron spits. Another bundle of spits together with Laconian pottery
was discovered during the excavation of Artemis Orthia temple at Sparta (7th century
BC) and the sanctuary in the Megalopoli street in the same town had three obeliskoi in a
THE BLACKSMITH’S STARTING STOCK 39
corner. In the temple of Hera Limenia at Perachora iron spits (of rounded section) were
also found both inside and outside the temple district; these were dated to between 750
- 650 BC. Iron spits are mentioned in reports on German excavations in the Heraion of
Olympia. Written evidence also exists for these deposits usually related to the votive
gifts of the temples. For example in the temple of Hera Limenia an inscription mentioned
a votive gift of 6 obeloi i.e. 1 drachme. A huge votive gift of obeliskoi was recorded at
Delphi. Herodotus and other authors refer to a hetaera named Rhodopis of Naukratis who
offered one tenth of her large fortune in spit currency to the temple at Delphi. This votive
gift was kept for many years in the shrine of the people of Chios. Inscriptions published
from the temple of Hera (Boeotia) proclaim votive gifts from the inhabitants of Thespiai:
3 drachms in Syphai, and 2 in Creusis. Several kinds of stylized currency are mentioned
in the Laws of Gortyna and in other inscriptions from this town in Crete, such as tripods,
obelisks etc. (7th - 4th centuries BC). Epaminondas of Thebes, a famous general and
statesman (4th century BC) died, according to Plutarch (Fabius 20.7) in utter poverty;
only one obelos was found in his house after his death.
Sometimes iron spits were used as a part of grave furnishings. In addition to the finds
from Fortetsa two graves were noted at Argos (8th century BC); in one of these a warrior
was buried with his armour and with some artefacts including two double-axes, two long
fire-dogs, and 12 obeloi. Another Argive tomb contained 6 items. The spits from these
graves were of a different type from those in the Heraeum; one end was hammered into a
circular plate (Fig. 16: 5).
Iron currency circulated in various regions of Greece for a long time, in later periods
contemporaneously with silver coins. They were used in internal commerce and trade.
In Sparta, iron spits survived up to the 3rd century BC which may be the origin of the
Lycurgian iron money legend of Sparta. It is interesting that in addition to these stylized
iron currencies also iron coins known as sidareoi (Fig. 16: 1) were also struck. They are
known from Arcadia, Tegea, Heraea, Thebes and Phokis (4th century BC). The sidareoi
of Byzantium were also mentioned by ancient authors.
The existence of an iron currency in ancient Greece proves that iron, although a com-
mon metal was still a valuable material. To a similar stage of development came later in
the early Middle Ages in European regions which were not imfluenced by the civilization
of Rome where the large-scale iron production resulted in application of other kinds of
semi-products, ingots and bars.

Roman ingots
Roman iron ingots were discovered in enormous numbers in Imperial provinces. They vary
in shape, dimensions and weight, and, in general, do not involve any stylized implements
and were of a very different quality: wrought iron, hard carbon steel often affixed with
punchmarks of the producers.

Bipyramidal bars

Compared with the heavy Celtic bipyramidal ingots the Roman pieces are relatively rare
and represent what can be called a miniaturized version. A typical example comes from
Martinsberg near Andernach (south of Bonn, W Germany, Fig. 17: 1); it is only 9.2cm
long, weighs 78g and is punchmarked with the name of C.RVBELLI (Fig. 17: 17) on all
40 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
of the four sides (Gaitzsch 1978, fig. 14). A piece from Magdalensberg, Austria, is punch-
marked with the name ORANI (53.3cm, 94g, Fig. 17: 2) and is dated to the Tiberian
time. It belongs to a set of similar unstamped pieces (9cm to 12.5cm, 75 - 118g, Fig. 17:
3) which are, however, called ‘Pfriems’, i.e. punches (Dolenz 1998, 219 - 222, pl. 79: W
364 to 368). Naumann (1964, 501 - 502, figs 13 and 14) and Schaaber (1964, 505, fig. 5)
called them Doppelstachel (bipointed thorns) and found two of them to have been made
of hypereutectoid steel (1.3 - 1.43% C, pearlite with cementite network) and another one
showed 0.36 - 0.64% C; they were inclined to accept a form of expensive trade material
or working pieces.

Block and rod ingots

When talking about miniaturized forms it is impossible to omit a small iron block (1st
century AD) from Oberwinterthur, ancient Vitudurum in Switzerland (Fig. 17: 4). It
is only 4.7cm long and square-sectioned (1cm x 1cm x 1.7cm, weight 55g. A stamped
punchmark reads [T]GC (Fig. 17: 18). It is not of steel but of relatively pure, in general
ferritic iron with low Mn and P contents (metallography by Pleiner 1996b, 195 - 196, fig.
184; 221 - 226, figs 219 to 226). Apparently such material was produced by renowned
workshops and wanted by customers for making small objects of desired quality.
The bulk of the other Roman ingots are, of course, brick-shaped heavy blocks. An
early find comes from the late Republican legionary camp at Cáceres el Viejo in Spain: 3
block bars weighed 4.6 to 5. 6kg (Ulbert 1984, 132 - 136, pl. 29: 264 - 265). The weight
apparently played a role and some authors reckon with the possibility that it could have
been related to Roman pounds (slightly over 3kg). A brick-shaped iron (Fig. 17: 8) said
to be a half or third of the weigth of which was supposed to have been 12 or 18 Roman
pounds (Ginouvez et al. 1998, 184, figs 4 and 5) was found in a forge of a Roman villa
at Sauvian-La Domergue Hérault, S France, 5th century AD). The Roman smiths and
merchants used balances to weigh iron ingots as a funeral stele from Augst - Augusta
Raurica (Switzerland) clearly attests: 20 pieces have been treated to some extent 42 are
ready for treatmenteith the aid of stone weights (Doswald 1994, 343, fig. 7; a list of
other rectangular block bars is discussed on pp. 334 and 335, fig. 6; the work with
balances see also on the mosaic from Sousse (below). According to the list mentioned,
individual examples or sets of the ‘lingots carrés’ were found in Roman sites including
military stations, vici, and villas, sometimes in boxes containing other metallic objects:
the distribution covers not only territories north of Alps but also Britain and Wales. The
item from a villa at Hambacher Forst near Aachen in Germany weighed 10.95kg (24cm x
12cm x 5.5cm (quoted by Feugère and Serneels 1998, 256, after Rech).
Several words should be devoted to two sites, one early and the other late Roman. The
first is the Austrian Magdalensberg whose layers yielded so numerous sources elucidating
the early blacksmith’s work: Apart from the ‘punches’ or small bipyramidal bars men-
tioned above Dolenz describes several kinds of iron ingots and bars; the ingots involve
heavy blocks (18.6cm x 11.9cm, 5.7kg; 14. 3cm x 9cm, 3.85kg, Fig. 17: 5 and 6; see
Dolenz 1998, 231 - 236, pl. 87: B1; pl. 88: B2) or fragments (2.66kg, pl. 88: B3). The
latter has been metallographically examined showing many surface cracks and structure
of a hypereutectoid steel and many large slag inclusions (Straube 1996, 126 - 130, figs 41
to 43). Possibly this piece was discarded as not easily workable.
The late Roman ingots were found in the eastern Roman province Pannonia Superior
THE BLACKSMITH’S STARTING STOCK 41
(now Croatia): at Sisak, ancient Siscium and at Hrvatska Dubica (Fig. 17: 7). The
Sisak ingots (3rd - 5th centuries AD) from the ‘Mint’ site weigh 4 - 7kg (length 20 to
40cm, those from Dubica comprise 28 preserved items (total weight 118kg) from 97 pieces
uncovered in 1880 (total weight more than 400kg (see Durman 2002). Both latter sites
are situated near one of Rome’s leading iron producing centres (Sava and Japra regions).
The circulation of iron ingots took place not only on land but also by coastal cabotage.
Nearly twenty ship wrecks have been discovered bordering Mediterranean coast from
Spain and France to Sicily and bear witness to iron ingot transportation (Feugère and
Serneels 1998, 252 - 262, fig. 3, after L. Long and A. J. Parker). Most of them have been
submitted to research, especially those along the southern coast of France dating from
the 1st century BC up to the 5th century AD.
The most important amongst the wreck finds with loads containing trade iron is a
site in the shelf of the Rhône river estuary near Saintes-Maries-de-la-Mer. Feugère and
Serneels (o. c.) inform us of various data which are useful for presentation in an abridged
abstract. Eleven wrecks were identified and information can be derived from the cited
study from about 7 of them. Wrecks (labelled as SM) 3, 6, and 8 are dated to the 1st
century AD. The loads consisted of iron ingots and bars which were classified into 6 forms
or types (Fig. 17: 11 to 16): 1 - 3 are square sectioned (5cm x 5cm, 4cm x 4cm) rods
95cm to 125cm or 42cm to 71 or 30cm long (found in SM 1, SM 2 and SM 6. Among
the pieces of type 2 one weighed 2kg and another one 3.27kg which is exacly 10 Roman
pounds (0.327 kg each). This gave rise to metrological reflections. The ingots denoted as
form 4 are 28cm long (3.9 - 4kg) and a little oblong in section (4.5cm x 5cm) and were
found in wrecks SM 2 (about 22 pieces), SM 8 (479 pieces) and SM 9. Form 5 is similar
(27cm, 4cm x 5cm, found in SM6). The form of 6 bars differ in that they are rather
flat blocks 29cm long, 10cm x 3cm in section, weighing about 6kg. It follows that some
wrecks (SM 2 and SM 6) transported ingots of several forms/types whilst other (SM 8,
SM 9 and SM 10) were loaded with one type. This can be interpreted thus: various ingots
in one load means small-scale merchandizing between producers and users or customers
whilst the loads containing one form were directed to customers dealing with large-scale
business ( o. c, 262). Some of the bars from different wrecks bear partly legible, partly
illegible punchmarks relating to names LEPIDI and EROTIS (SM 2), MARI (SM 3), C.
RVTILI (Fig. 17: 19) and AP (SM 6), CICELLI (SM 9), ..LISES (SM 10). No data on
the metal quality of the stamped Sainte-Maries irons were published until now. Instead,
another piece which is eloquent in this respect is at our disposal: a flat ingot from a wreck
which sunk near Bonifacio in the Corsican strait has been stamped three times with the
smith’s name SATVRNIN[VS]; it was forged of hard carbon steel containing 0.9% to
l.0% C (pearlite in cementite cells); the dimensions: 23cm x 12.5cm x 6cm) Another, not
stamped bar was 23cm long and its cross-section was 5cm x 5cm; this showed ferritic iron
(Zwicker 1996). An ingot from the Palavas wreck, east of the Bouche de Rhône beared
a stamp of the HAEDVI and another one found in the wreck near Ben Afelı́, Spain, was
punchmarked with the word FERRO. Apparently, trade iron from different workshops
was transported by boat and further circulated through selling, buying and use. The
handling with ingots illustrates very well a 3rd century AD mosaic at Sousse, ancient
Hadrumetum in Tunisia, presenting the unloading of about 1m long ingots from a barque
and controlling them on a balance using (stone) weights (Feugère and Serneels 1998, 257,
figs 7 to 9).
The finds from other wrecks indicate, iron was forged in different ‘forms’ of ingot,
42 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 17: Roman ingots. Small bipoints: 1 Martinsberg, Germany, 11.2cm; 2 - 3


Magdalensberg, Austria, W367, 13.3cm; Miniature wrought iron bar: 4 Vitudurum-
Oberwinterthur, Switzerland, 4.7cm. Block ingots: 5 - 6 Magdalensberg, Austria, B3
and B2, 14.3cm x 9.5cm for B2 7 Hrvatska Dubica, Croatia, 18cm; 8 Sauvian-la Domer-
gue, S France, 8cm x 11cm; 9 - 10 Magdalensberg (9 B24 12.3cm, 10 B12 38cm). Rod
ingots from ship wreck at Saintes-Maries, S France (different scales): 11 form 1, ship
wreck SM 10; 12 form 2, SM6, 46 cm; 13 form 4, SM 9; 14 from 5, SM 6; 15 form 3, SM
6; 16 form 6, SM 6,30cm x 10cm. Punchmarks on 1, 4, and 11: 17 RUBELLI; 18 [T]CG;
18 RVTILI. 1 - 6, 9 - 16 early Roman, 7 - 8 Late Roman. 1 and 17 after Gaitzsch, 2 - 3,
5 - 6, 9 - 10 after Dolenz, 14 and 16 after Pleiner, 7 after Durman, 8 after Ginouvez, 11 -
16 after Feugère and Serneels.
THE BLACKSMITH’S STARTING STOCK 43
sometimes equivalent to those from Sainte-Maries, sometimes differing in weight and size
(Feugère and Serneels o. c., 255 - 256). As the loads have been transported they could
not have been regarded as negligible (Plemirio B, Sicily, 1 ton); in other cases in addition
to iron the ship carried other goods as well, oil or wine in sets of amphoras (e.g. Les
Sorres, coast of Spain). The traffic started some time about 100 BC (Bagaud B, Capo
Testa B in Sardinia, 1st centuy BC) and continued during the Roman period; a ship (the
wreck Mateille A) sank near Gruisson after AD 400.

Various flat bars

The finds from Roman sites also involve flat forgings of iron which are interpreted as
bars for further use in smithies. In the early Roman strata of Magdalensberg were un-
covered flat bands (e.g. B 12, 31cm long, 1.74 kg) and sets of triangular examples with
pointed tips, about 12cm long, 0.95kg (Fig.17: 9 and 10, see Dolenz 1998, pl. 98 and 99:
B 24 to 30). Fluzin (2002, 71 - 72, pl. III: 20) depicts a bar (21cm, 2.41kg) from Toufre-
ville in Calvados, France (Gallo-Roman). No details can be presented concerning the two
1st century AD bar hoards from a fortified settlement Niederzier-Hambach, Germany.
Blanks forged as stylized implements document, again, various forms of trade iron (and
possibly local means of exchange) in European countries having been not affected directly
by Roman civilization and which just have started their iron industries on a more exten-
sive scale.

Early medieval tool-shaped blanks


The custom of shaping iron blanks into stylized implements continued documenting that
various forms of trade iron could have been used, possibly in local circumstances, as a
local means of exchange. This was the situation in in early Scandinavia and subsequently
in central and eastern Europe.

Spade-shaped bars

The excavations of a Vendel period centre (6th - 7th centuries AD) at Helgö in cen-
tral Sweden initiated more detailed studies on iron ‘currency bars’. Three iron objects
resembling spades or socketed ploughshare staffs were found within building 2 (Fig. 18:
1 and 2). They have many counterparts in the above-mentioned part of the country
(southern Norrland in Sweden with some more isolated finds in Gotland, Åland and at
Trøndelag in Norway). They were forged in groups of 20cm - 28cm, 28cm - 34cm, and
34cm - 44cm in length (weights 0.154 - 0.35 kg, 0.35 - 1 kg, 1 - 1.65 kg). The tips (and
sometimes the sides) of narrow blades were often folded over to achieve the desired dimen-
sions. Numerous hoards of these bars have been recorded unfortunately displaced during
agricultural work, e.g. Offerdal (60 items), Sundsvall (15), Biskopskulla (21) Arbrå(about
104 in 1880), Hökbäck (61 plus 79, deposited in 5 - 6 layers), Valbo (58) - in sum, about
1500 pieces from 92 sites. The dating is assumed to range from the Vendel to Viking peri-
ods. The objects in no way represent normal tools and have been interpreted as currency
bars (Hallinder and Haglund 1978, 30 - 37, 49 - 54, figs 1 to 8).
44 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 18: Swedish early medieval iron blanks. Spade-shaped currency bars 1 - 2 Helgö,
Vendel period, about 35cm - 30cm. Rod-shaped currency bars: 3 - 4 Skedstad, Öland
(3 the hoard, 4 single, double and triple bars. Scythe-shaped currency bar: 5 Fagerhult,
Småland. After Hallinder and Haglund.

Scythe- and rod-shaped bars

The same verdict applies in the case of elongated iron bars of triangular cross-section
and slanting tips (Fig. 18: 5). The length varies between 47cm and 79cm (the bulk being
52cm - 72cm long) and the weight in the preserved state between 0.15 - 0.6 kg. The
distribution mainly covers Småland in southern Sweden. 22 sites have been registered
certain of which were definitely hoards although in many cases they have been removed
or only superficially documented. At Furingstad, Östergötland, 5 hoards are reported to
have been discovered in 1928 in a sand pit, containing 300 - 500 bars. Other important
sites are Ventlinge (46 pieces plus 50 fragments), Törnbotten (44 under a slab of lime-
stone), both in Öland, Rumhult, Fagerhult (340 - 400), Frödinge (120) and Ljungby (40 -
50) in Småland (Fig. 18: 5), 3 examples among the stray finds were bent and folded like
tweezers. The metallography of one of the bars from Fagerhult showed it had been welded
together from 2 - 3 bands (0.1 to 0.8% C) with prevailing low carbon steel (Hallinder and
Haglund 1978, 39, figs 4 and 5, pp. 43 - 45, 53 - 54, 57).
In connection with the scythe currency bars attention has to be paid to the so-called
rod-shaped bars which are thinner (square-sectioned) and, according to K. Haglund, have
slanting tips like the scythe-shaped pieces. Otherwise they should be held to be specific
form of the bar which is remarkable by the fact that some items have been bent or
welded into tweezer form (the letter V like the above mentioned scythe bars) but, on rare
THE BLACKSMITH’S STARTING STOCK 45
occasions, twice bent forming the letter N (Fig. 18: 3 - 4). The hoard from Skedstad,
Öland, contained 560 bars and 2 fragmnets smoothly arranged in a pit 65cm - 65cm. It
consisted of 98 double bars and 2 triple N-pieces. The length is about 50cm (Hallinder
and Haglund, o. c.39, figs 6 to 11). The weight has been influenced by corrosion (0.052
- 0.148 kg). Among the other sites from which sets or ‘bundles’ have been reported the
following examples have to be pointed out: Furuby, 39 pieces; Klöckeberga, Algustrum in
Öland, Eketorp fortified site with 15 items from the site area, Gräsgård with 200 items
found in a burial mound found in 1825). They have been dated to the 5th to 7th century
AD (the site of Helgö, Vendel period 6th - 7th centuries, yielded two bars of this kind (o.
c. 52).

Respecting the Swedish authors’ terminology in English another group of rod-shaped
irons has to be mentioned which were also found at Helgö and have been defined as
rod-shaped blanks (not rod-shaped currency). They are considered as a virtual working
material bearing traces of chopping and cutting (Hallinder and Tomtlund 1978, figs 1 to
5); 38 have been metallographically examined (Tomtlund, ibidem, 77 - 80). Four different
groups ranging from thicker square-sectioned tod slender and often round-sectioned were
presented; the latter were usually of hard steel (pearlite, pearlite-and-cementite up to 0.8
- 1.2% C). However, the carbon content was heterogeneous. Various kinds of material
were obviously used for making different sorts of artefacts.

Plough iron

Here we have to consider the ploughshare blanks again. Some of them were found in
Sweden (Råaby, three ploughshares inserted socket-into-socket, see Nihlén 1939). In early
Russia ‘lemeshnoye zhelezo’ circulated up to the 15th century and is noted in written
records together with ‘rods’and ‘discs’ (Kolchin 1953, 45).

Axe-shaped bars

Here we are dealing with slender iron objects equipped with a minute shaft hole at one
end, an elongated body and a rounded blade. It is interesting that these stylized axes
appeared in quantities in two different, mutually isolated geographical areas, in Scandi-
navia, above all in Norway and, on the other hand, in territories colonized by the western
Slavs in central Europe (Moravia, Slovakia and southern Poland).

Scandinavian axe bars.

In Scandinavia the objects in question began to circulate at a relatively early date. In


Snorup, the well-known iron-smelting site (Voss 1966) a bar hoard was found in a long-
house which contained 6 massive iron ingots (0.025 - 0.7% C interpreted as material of
local production) and 200 axe-shaped bars in two sizes: 100 items were 26cm - 30cm long
and another 100 16cm - 20cm weighing 0.12 - 16kg and 0.025 - 0.03kg respectively. These
axe-shaped products (Fig. 19: 1 and 2) have been interpreted as imports from Norway
(Høst-Madsen and Buchwald 1999). The finds province in Gudbrandsdalen in eastern
Norway yielded the bulk of these bars which ocurred in two groups of similar dimensions
46 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
(Svane 1991). About 110 hoards the largest of them producing 400 - 600 pieces have
been roughly dated to the latest Migration period up to the Early Viking Age (6th -
8th/9th centuries AD. The information is rather fragmentary. The structures of the bars
were mostly pearlitic and welding seams have been observed; the blades are more pure
in terms of the amount of slag inclusions (o. c, 31 - 32). No chopping marks have been
reported (Fig. 19: 3 - 6).
When dealing with the axe-shaped bars we cannot avoid, for ethnic reasons, to insert
an interluding section devoted to another kind of iron object preceding and contemporary
with the mass appearance of axe bars in early Slavic cultures.

Slavic saucer-shaped disks.

Not only implements but also domestic utensils might have served as models for forg-
ing blanks. Here we should mention the so-called ‘Silesian iron discs’ of about 20cm in
diameter and 5mm - 20mm thickness and usually severely corroded (Bubenı́k 1972; Dostál
1983; Pleiner 2000a). Nevertheless, even in this state they are also relatively heavy (0.6
- 0.8kg, the heaviest being 1.72kg). In certain cases they were pressed together from 2
- 3 layers of iron sheet. Their geographical spread covers Bohemia, Silesia and southern
Poland, Bohemia, Moravia and Slovakia, about 60 sites have been registered including
hoards of discs alone (Silesia) or hoards of discs together with other iron objects. Their
function has been the subject of lively discussion (roasting pans, fumigation plates, iron
blanks, currency bars). The earliest disc comes from a pit in the early 6th century AD
settlement at Březno in NW Bohemia (Fig. 19: 15) but the bulk is a little later, occurring
up to the 9th century. An important point is that that this kind of disc was found in an
accumulation of iron objects together with several axe-shaped bars which are discussed
below: in hut feature 21/V-XII in the Great Moravian centre of Pohansko (S Moravia).

Slavic axe-shaped bars.

The Slavs colonized the central parts of Europe in the late 6th century AD and about
two three centuries later, in the period of Great Moravia (Megalé Moravia after Porfyro-
genetos) they intensified their production of iron (Pleiner 2000c, 190 - 193, 276). As a
consequence, forms of trade and barter iron entered into circulation, above all the Slavic
axiform bars. Their typological description includes sub-types (Pleiner 196l; 1981): those
with a miniaturized shaft hole with lateral supporting pointed plates (similar to early
Slavic battle axes and other objects), then, those with a rough perforation instead of the
shaft hole; those with supporting tips fore and aft the shaft hole and, rarely, those with
hooked tips instead of a shaft hole, Fig. 19: 8. The lengths and overall dimensions which
fall into four categories are very important: large (about 40cm - 47cm long), medium
(21cm - 28cm), small (15cm - 17cm), and miniature (4cm - 9cm), see Fig. 19: 8. These
striking dimensional categories may be meaningful in terms of value and would argue in
favour of the currency possibility. In many cases small plates and triangles were cut off
(Pleiner 1961, figs 9 to 11), apparently in order to get small pieces of material for further
handling, e.g. for secondary carburization, forming small objects etc. (Fig. 19: 9 - 14).
Some pieces were deformed as a whole. The role of a means of exchange would have been
thoroughly mixed with that of working material.
More than 20 specimens from Moravian and Slovakian sites have been metallograph-
THE BLACKSMITH’S STARTING STOCK 47

Figure 19: Scandinavian and western Slavic axe-shaped bars and accompanying ingots.
Scandinavia: 1 - 2 Snorup, Jutland, two sizes of axe bars, 25cm and 13cm and a massive
ingot, 13cm; 3, 5 Krokerud, Norway, both about 25cm; 4 Norway, about 12cm; 6 Vestre
Toten, Norway, about 8cm; 7 Aardal, Sweden, hooked ingot, 16cm. Slavic axe bars: 8
four categories: large, about 45cm; medium 25cm - 30cm; small about 15cm; miniature
4cm - 9cm; 9 - 14 axe bars with cutt-off parts: 9 Mikulčice, Moravia, 10 - 13 Hrádok,
Slovakia, 14 Nejdek, Moravia, 15 disc-shaped iron plate from Březno, Bohemia, dia. 20cm,
16 - 18 examples of miniature bars (16 Mikulčice, 4cm, 17 Brno-Lı́šeň, Moravia, 8.5cm,
18 Staré Město ‘U Vı́ta, hoard, 9cm); 19 Vel’ký Klı́ž, Slovakia, about 25cm. 1 - 2 after
Høst-Madsen and Buchwald, 3 - 4 after Svane, 5 - 18 after Pleiner and various authors,
19 after Bialeková.
48 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
ically examined (Pleiner 1961; idem, unpublished, archive of the Arch. Inst. Prague;
Bialeková et al. 1999). The picture shows not unexpected outlines: mostly ferritic
wrought iron or ferritic-and pearlitic low carbon steel with an uneven carbon content;
faggotting and piling has been not frequently observed.
Within the Moravian and Slovakian finds province the bars occur in settlements (espe-
cially in leading centres), sporadically in graves (Žlkovce, 3 items in a grave), and the bulk
is known from numerous hoards ranging from tens up to hundreds of pieces; again, they
came to light in prominent fortified sites. Examples from Moravia: Brno-Lı́šeň - a hoard
in a hut, 13 pieces; Staré Město - stray finds and a hoard with 11 bars in bundles, 3 - 4
items each (Ohlhaver 1939; Pleiner 1961, 409 - 412; At Staré Město ‘U Vı́ta’ a hoard of 32
miniature bars and 20 fragments was deposited in a smithy (Galuška 1992, 145 - 149, figs
12 to 14), presumably used as scrap for recycling; Nejdek - a hoard with 21 axe bars plus
fragments and other iron objects; Mikulčice - stray finds and a hoard in the presbytery
of church 8. Slovakia: The most important site is the fort at Pobedim, a trading post in
the Váh river valley - 22 hoards of axe bars were found, about 2200 pieces in total (e.g.
Bialeková 1981, fig. 3); Hrádok near Trenčı́n on the same trading passage - a hoard with
originally 400 bars in bundles of 20 items each (Pleiner 1961, 415 - 418, 427, fig. 16);
Vel’ký Klı́ž - hammer, agricultural implements and 40 axe-shaped bars, both complete
and fragmentary (Bialeková et al. 1999. Fig. 19: 15 - 19)

The territory of Southern Poland north of the Carpathians appears as a region with
interesting finds of axe bars that have been modified in their shape (Zaitz 1990). The
number of sites is limited but their importance is inherent in the existence of a huge
deposit discovered in Cracow (see below).
Leaving aside a piece from Nowa Huta-Mogila, the hoards of Radymno (12 pieces)
and Zawada Lanckorońska have to be mentioned. The latter consisted of 10 bars (about
30cm long, 0.45 - 0.6kg, shorter and lighter pieces about 15cm, 0.4kg). The necks of
the bars are rounded and perforated. Three examples were submitted to metallographic
studies by Piaskowski (1956) which revealed ferritic and ferritic-pearlitic structures and
a considerable phosphorus content of 0.5 - 0.77% P. At Piotrawin a hoard of 12 well
elaborated axe-shaped bars with lateral triangular supporting trunnions at the shaft hole
was found (weight 1.5 - 2.2kg, see Sulowska 1977). Fig 20: 6 - 7.
As regards the volume of the buried iron and the number of bars the top position
belongs to the ”bar hoard of the 20th century” which has been uncovered at Cracow-
Kanonycza street No 13. This place was in reach of the centre of the Slavic Wiślanie tribe
which was the fort of Wawel. During structural works in the house cellar a rectangular
pit (210cm x 110cm, 100cm deep) was excavated. The walls were timbered with oak
and fir planks. Inside there were 27 layers of axe-shaped bars in 3 columns (Fig. 20:
top) consisting of 150 bundles (3 to 15 pieces each, mostly numbering 7 - 9 bars, tied
with organic basts). The total amount comprised 4212 bars with a total weight of 3.63
tonnes. The top layer was sealed with sand sintered by corrosion products into a coherent
mass. The bars occured in two main subtypes: the minority is represented by objects
with a rounded neck as briefly described in the case of Zawada (above), the bulk involves
pieces with a clearly formed shaft hole and with minute vertically protruding trunnions
on the neck. All of them belong to the larger category (Fig. 20: 1 - 5) and the length
varies between 32cm and 37cm and the weigth between 0.5kg - 1,7kg. About 35% of
THE BLACKSMITH’S STARTING STOCK 49

Figure 20: Polish axe-shaped bars. Top: Cracow-Kanonycza street No 13, general view
of the hoard in a pit (photograph Jan Barsı́k); 1 - 5 examples from the hoard; 6 Zawada
Lanckorońska; 7 Piotrawin, 8 Čebovce, Slovakia. After Zaitz.
50 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
the objects show welds: either as longitudinal piling traces or transveral butt-welding.
Comments concerning the metallographic test carried out on 15 examples: The brittle
metal contained a high phosphorus content of 0.1 to 0.99% P and mostly consisted of ferrite
with a varying grain size. Lamination could be observed in four of the examined bars
(ferrite-and-pearlite, 0.2 - 0. 3% C; white lines with arsenic and phosphorus segregations).
Broken slag inclusions indicate the final forging took place at lower temperatures.
The reason for the deposition of the huge Cracow bar hoard is difficult to understand.
There are no clear signs of a votive sacrifice. It may be it was really treasure belonging to
the rulers of the Wiślanie tribe accumulated from tributes and taxes in iron from various
groups of people and held in reserve.
The circulation of blanks in the form of stylized implements is a remarkable phe-
nomenon in the economy of materials. It signalizes a transitional period at the dawn
of the large-scale production of iron in different times and areas during which the role
of these intermediate products appeared to swing between a starting material for smiths
and object of trade up to a means of exchange and even premonetary currency (in Ger-
man Gerätgeld). This stage concerned ancient Greece (8th to 4th century BC), the early
Celtic world (3rd to 1st centuries BC), the Scandinavian earliest Middle Ages (from the
6th to about 9th centuries AD) and the western early Slavic cultures (9th century AD).
(By the way, exactly the same role was atributed to African iron currencies in the 19th
- 20th centuries, see Pleiner 1961, 436, 438 - 439, fig. 23, with references). Within the
civilization of ancient Rome this role is indistinct or veiled in the early period; the later
mass production of iron and steel for trade is manifested in the marginal territories of
advanced Roman provinces.

Other medieval bars

During the Viking Ages some workshops in southern Sweden produced a flat kind of
propeller-shaped bar two examples of which were first recognized in the hoard box at
Mästermyr on the island of Gotland. They were named the Mästermyr type (Fig. 21:
1 - 3, see Hallinder ind Hallinder and Haglund 1978, 45 - 47). The number is not high.
Apart of another two bars known from a 1880 find at Rabelövsjön three turned out at
Nosaby (Skåne). Two bars were found in layers of the town centre at the fortified site
of Haithabu (S Jutland) and about twenty as a hoard (Fig. 21: 2), apparently imported
ware (Müller-Wille 1980). The bar labelled as No 3 was 45cm long, 4.5cm wide and 0,6cm
thick weighing 0.62kg. The metallographic analysis of numbers 1 and 2 revealed ferritic
and ferritic-pearlitic phases (0.2 - 0.4% C). The carbon content was spread unevenly
and many welds show that the bodies were joined together from several bands. There
is a welded-on steel splinter on the side of bar 1 (Thomsen 1971). (In addition to the
Mästermyr bars three heavier oval iron blocks were found one of which measured 15cm
x 5.5cm x 3cm, weighed 1.45kg and had been welded from several pieces of wrought iron
and hard steel the average carbon content being about 0.6% C but a small area displayed
local enrichment with 4.3% C and a ledeburite structure, see Thomsen, o.c, figs 2, 4, 5, 8
to 110).
The earlier and later centuries of the Middle Ages witnessed the dying out of the prac-
tice of depositing iron blanks in hoards. Therefore, there is a lack of archaeological evi-
dence concerning the starting material for smiths. A small rectangular bar (Sou/31/1084,
THE BLACKSMITH’S STARTING STOCK 51

Figure 21: Medieval iron bars. Mästermyr type (9th century AD): 1 Mästermyr, Gotland,
about 49cm; 2 Haithabu, Jutland, about 45cm; 3 Nosaby, Skåne, about 40cm. High
Middle Ages: 4 Piekary hillfort, Poland, about 20cm; 5 - 7 Salzach river ship wreck near
Bergheim, Austria, 15th/16th century AD (split bloom, derived billet, block ingot. 1 - 3
after Müller-Wille, 4 after Piaskowski, 5 - 7 after Feldinger.

20mm x 9mm, 25g) with one overlapped end, from Hamwic near Southampton, S Eng-
land, has been examined revealing a high carbon content of about 1% C, as did another
fragment (Sou/31/2261, 6cm a 1cm, 39g) which had been embedded in a glassy iron-poor
slag and consisted of cast iron with 3.5 - 4% C. A larger Hamwic billet (Sou/31/2780,
10cm x 8cm x 4cm, 0.6 kg) was fine pearlitic and bainitic with traces of austenitic grains
containing 0.8 to 1.2% C and being decarburized on the surface down to 0.2% C (Mack
et al. 2000, 90 - 91, figs 1 an 2, and 93 - 94, figs 9 to 11). The authors of the exami-
nation are ready to classify many objects from Hamwic, mostly steel knives, as derived
from from decarburized cast steel, either imported or smelted locally in low furnaces in a
rather unknown process. The question deserve further study within the framework of the
appearance of early cast iron in other archaeological contexts.
At Biskupin and above all at Piekary (Fig. 21: 4) in Poland short rods (5cm to 20cm)
of low carbon phosphoric iron came to light (0.15 - 0.94% P, see Pleiner 1962, 172 with
references to Kapitańczyk and Piakowski). Written records mention the so-called šı́ny,
Schienen, literary rails, as semi-product forged in hammer-mills.
A specific form was located with iron split bloms in the sunken ship wreck in the Salzach
river in Austria (Fig. 21: 5 - 7). They resemble ‘butterflies’ but are in fact spread-out
blooms mentioned (Feldinger 1990, 150, fig. 118: right). The date is 15th/16th century.
52 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Little information can be extracted from metallographic studies on several of the hetero-
geneously or highly carburized rectangular iron blocks from Belarus (Mysli, Taymanova,
Abidini). Neither dimensions and weight nor chronological positions (postmedieval?) and
find circumstances are described (see Gurin 2001).
As has been shown, the quality of ingots and blanks was influenced by the hetero-
genelous presence of carbon. Of the samples so far examined, the minority has been
made up of hardenable steel.
The question remains by what method the smith distinguished and separated the
required hard steel from the billets when undertaking, in certain advanced regions and
chronological periods, the welding togeher of iron and steel in the construction of tools and
weapons. It will be mentioned elsewhere that the ancient smiths were able to empirically
guess the nature of the material they had to work but what escapes us is how they treated
the uneven qualities of ingots and billets.
The wide range of shapes of iron starting materials and blanks shows that they them-
selves were often very complicated iron forgings made in a sequence of operations like
cutting, drawing dawn, splitting, bending, rolling and piercing. Therefore, the forming of
cold and hot iron has to be treated in detail.
Chapter V

FORGING OPERATIONS

Smith’s operations during which iron and steel were transformed into usable things,
tools, weapons, utensils and mountings include in principle shaping and heat treatment.
The following lines attempt to describe, in an abridged form, the individual phases and
techniques known from modern surviving hand smithing and can be attested by the study
of archaeological finds.

The forming

The most important way to shape iron and steel up to 1.7% C is hot forming - the working
of the material above 723 ◦ C when metal with 0.8% C (eutectic point) reaches the plastic
austenitic phase (equilibrium diagrams Fe - C and Fe - Fe3 C). The lower the carbon
content, the higher the temperature is needed for the transformation of iron into pasty
austenite. Carbon poor metal (0.02 %C) must be heated over 900 ◦ C to be hot worked.
The presence of carbon in the solid state of steel is visible on polished and etched samples
in form of different crystalline phases resulting after the slow cooling of austenite: 0.03%
to 0.8 % C various proportions of ferrite and pearlite (consisting of minute lamellae of
ferrite and Fe3 C cementite), at 0.8% pearlite, at 0.81% to 1.7% pearlite and cementite.
Naturally, the smith of the past did not know either the the metallographic phases nor
the temperature of the metal worked. They used their senses. By his ear he empirically
tested the sound of two metal pieces hit one against the other - harder steels clank
keenly and, when hanging, sound for a longer while. He could controll the sparks of a
struck sample - wrought iron developes droplet-shaped sparks whilst harder steel produces
cracker-shaped sparks. He could try to recognize the hardness by the use of a file or by
bending test stripes - with a wrought iron bar (up to 4mm thick, containing up to about
0.15% C) a full can be made but steel with about 0.2% C shows cracks on the outer
side after this operation; a steel band with about 0.3% C can be bent to a 40◦ angle but
cracks appear again with increasing carbon content (over 0.35% C). Other tests can be
made with glowing metal and the early masters used many ways of experience how to
gain experience how to recognize softer and hard, bad or good iron.
The smith did not have thermometers available but their eye was able to estimate the
temperature by means of the colour of the glowing metal. It is known that iron/steel
starts to glow when a brownish colour begins to appear (530◦ C in the centigrade scale)
and the smith could start the forging because this mitigates internal tensions and allows
the recrystalization of grains deformed by previous cold hammering. Hard steels can be
worked at dark red to bright red glowing (between ca 650-830◦ C) because the danger of
oxidation and even burning arises at higher temperatures. Soft steels and carbon-poor
wrought iron display orange up to white glowing colours (880-1250◦ C).
The estimating of temperature based on glowing colours is, of course, largely subjective
and an appropriate darkness is favourable.
The metal has to be heated evenly - areas with varying temperature behave differently
during the heating. Large heated blocks may prepare difficulties because the inside might
54 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

C glowing colour

550 beginning brown


580 - 650 brownish red
650 - 750 dark red
750 - 780 dark cherry
780 - 800 cherry
800 - 830 bright cherry
830 - 880 bright red
880 - 1050 orange
1050 - 1150 dark yellow
1150 - 1250 bright yellow
over 1250 white
Table 1: Colours of heated iron.

be not hot and plastic enough whilst the surface begins to burn. The final forging tem-
perature is also of importance. The forging process should be finished at about 900◦ C
(orange) when working low carbon steel, and about 750-800◦ C (cherry red) in case of
harder steels. This prevents the crystalline grain growth. The smith has to take account
with this in terms of rapid cooling of the material and adjust the pace of his work. The
cooling could cause defects especially with regard to carbon steel. It should be cooled
slowly and gently. The smiths used to cover them up in a dry ash heap. Each heating
means a loss of metal. Therefore, experienced smiths work quickly not to repat surplus
heatings.

Principal shaping operations


A metal stock heated in the right way is ready to be manipulated with tongs and formed or
shaped by strokes of a hammer with the aim of producing produce simple forgings (struc-
tural elements, fittings) as well as sophistically constructed and shaped objects (tools,
weapons, ornaments) sometimes of an artistic value. Simple forging can be made by
using just some of the hammer operations whilst the complicated objects require a com-
bination and logical sequence of individual acts. The basic operations of shaping involve
cutting, upsetting, drawing-down, shouldering, bending and rolling, twisting, splitting,
piercing, welding, (possible swaging), planishing and sharping.

Cutting

A piece of iron/steel which is intended to be tranformed to various artefacts must be


divided from larger block or longer rods when hot: bars and billets. This is done on
an anvil edge eiher with a hand chisel or handled shafted implement (Fig. 22: A). The
smith should avoid a direct blow on the anvil edge because he would destroy both the
chisel and the anvil. The use of a low cutter set into an anvil hole is doubtful when the
archaeological finds are reviewed. Thinner wire or metal sheet could have been cut when
cold by chisels with the cutting-edge angle of about 60◦ .
FORGING OPERATIONS 55
Upsetting

This operation means increasing of the size or diameter of the worked piece by strokes
directed in vertical position. The English term migt be confused with the German word
Absetzen which means a completely different operation (see shouldering). Upsetting is
employed when the shank of the object has to be thickened. In the majority of cases
upsetting was practiced during the forming of the nail and rivet heads or handles and hilt
tips and other ends (22: B). The upset ends of rods could have been useful in preparation
of other parts for welding-in.

Drawing-down

This operation leads to the decreasing change of dimensions of the worked piece which
can be drawn-down to be thinned and lengthened or widened (Fig. 22: C-D). The first
can lead up to forging of wire, the other to the making of metal sheet. Poles, shanks,
blades and their handles can be made by the prolonged use of the peene of a hammer
and/or the edge of the anvil and bottom fullers. The red hot piece is fullered transver-
sally and the grooves have to be planished with the hammer face. The metal cools down
rapidly and the master has to work quickly. In the process of widening or spreading the
hammer peene strikes are parallel to the object axis and begin in the middle of the piece.
The cutting-edges of axes and blades of every kind, fittings and metal sheets are results
of widening which have been modified by the drawing-down process. This technique was
required when preparing the semiproduct for further operations, e.g. the splitting. Again,
the traces had to be, again, planished or smoothed with a set hammer.

Shouldering

In shouldering an abrupt change of the cross-section is performed. Step-like features


appear on forgings (Fig. 23: E). Endings of bars or rods are shouldered on one side or on
all sides and the thinner tips can be drowned-down by lengthening, facetted or rounded.
Handles or blades and cutting parts of axeheads etc. are results of shouldering. An anvil
working face edge and set hammer were used to perform this operation.

Bending and rolling

It was possible to bend thinner rods and wire in the cold state. More solid pieces were,
however, bent when red-hot. Bending in an angle and in a curvature has to be distin-
guished: the performation technique is different (Fig. 23: F). In the first case the anvil
edge and a hammer have to bee used. In this way hooks, clamps and keys have been
made since the La Tène period. Anvil horns were suitable when making arched bends.
Anyway, the masters of the Hallstatt period were able to produce splendid bent things like
complicated horse bits. The La Tène period smiths produced a wide palette of hinges,
hooks, chains with various shapes of links, springs of shears, overlapped shaft holes of
tools and a lot of work concerned with the production of ornaments, fibulae, rings etc.
Rolling-up is a kind of bending widened plates or sheets and was especially applied in
shaping the sockets of implements and weapons (e.g. spear- and lanceheads). An edged
bending was typical for forming sockets of the La Tène period socketed axeheads.
56 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 22: Forging operations. A Cutting. B Upsetting: 1 and 4 La Tène, Switzerland; 2


and 3 Uherský Brod, Moravia, early medieval; 5 Jarohněvice, Moravia, 9th century AD;
C - D drawing down: prolonging and widening: 6 Količı́n, Moravia, 9th century AD;
7 Semice, Bohemia, 10th century AD; 8 Ivanovice, Moravia, 9th century AD. Various
authors.
FORGING OPERATIONS 57

Figure 23: Forging operations. E Shouldering: 1 and 2 La Tène, Switzerland; 3 Steins-


burg, Germany, Celtic oppidum; 4 Moravský Ján, Slovakia, early medieval (welding comes
into question); 5 Čáslav Hrádek, Bohemia, early medieval; 6 Žitavská Tôň, Slovakia, early
medieval. F Bending and rolling: 7 Lovosice, Bohemia, Hallsatt period; 8 Stradonice op-
pidum, Bohemia; 9 Ponětovice, Moravia, La Tène period; 10 Ostprignitz, Germany; 11
Haithabu, N Germany, Viking period; 12 Staré Hradisko oppidum, Moravia; 13 La Tène,
Switzerland; 14 Dobřichov-Pičhora, Bohemia, Romano-Barbarian. Various authors, see
Pleiner 1962.
58 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Twisting

Twisting of rods and shanks was often used in antiquity, European prehistory and later
periods. The piece which has to be twisted must be fixed in a firm slot and the turing
along the axis is performed with tongs. The twisting hindered easy bending or could
have had a decorative purpose as well (Fig. 24: G). Medieval twist drills were made by
twisting steel and sharpening the cutting-edge.

Splitting

Ends of workpieces which have been partly cut or split by an obliquely held chisel can
be submitted to further procedure, e.g to bending (forks, dagger hilts, arrowheads, lynch
pins (Fig. 24: H). A specific splitting operation is the preparation of raw teeth. Hot
splitting requires a hammer, tongs abd chisel.

Piercing

A quicker and most advantageous method how making holes in iron objects is piercing.
Functional piercing was widely applied when providing the tools and weapons with shaft
holes: with rectangular chisel or round sectioned drifts and punches: hammers, fullers and
other implements. Circular openings have appeared since the earliest times: Scythian
shaft hammer-axes, Romano-Barbarian and Slavic early medieval axeheads, steels for
striking fire. Any hinged instrument must have been pierced with round-sectioned drift
(tongs, medieval hinged scissors, metal sheet shears). In the case of more massive arte-
facts the place indebted to be pierced would be heated to a yellow heat (about 1000◦ C
and upset to get larger dimensions. Already in the Hallstatt period flanged hilts of iron
swords and daggers were perforated to be equipped with rivets in the handle parts as
well as iron wheel tyres which were fixed with nails. Then, various mountings and fit-
tings were equipped with holes and openings and small holes to be fitted on to any type
of material. Ornamental perforation was more sporadic (e.g. Romano-Barbarian shears
from a grave at Dobřichov-Pičhora in Bohemia, La Tène period or Germanic lanceheads
with crescent holes). The operation of piercing must have been performed on an anvil
work face equipped with a hole (e.g. nail hole). For exemples of pierced and perforated
archaeological finds see Fig. 25: I.

Forge welding

The purpose of forge or hammer welding (in German Feuerschweissen, in Russian kuznech-
naya svar’ka) was the quick inseparable joining of metals (Fig. 26: J). Apart of iron other
metals are weldable as well: silver, gold, lead (Tylecote 1986, 109; 1987, 228, 242, fig.
6.36). However, forge welding played a most important role in the history of iron technol-
ogy. It comprised both joining of individual pieces and endings of iron artefacts (tyres,
rings, sockets). The smith used tongs for grasping the glowing metal, the anvil as a hard
support and hammer to develop strokes on heated plastic metal causing the penetration
of austenitic grains. Since the welded surface had to be clean and free from hammer scale
and slag film, a flux was necessary. Modern hand smiths use borax (sodium borate), in
the past it was mostly fine sand dissolving the hammer scale and producing the liquid
FORGING OPERATIONS 59
slag (see Chapter VII).
Four principal kinds of forge welding were practiced by early masters. (1) The joining
of more or less homogeneous pieces of low carbon wrought iron to larger blocks under
raised heat (yellow or white glowing colour, about 1100 - 1200 ◦ C). Thus, the La Tène
period bi-pyramidal ingots were welded together from several blooms which has been
metallographicaly attested on a find from Kaisheim (Rädeker and Naumann 196l); the
Celtic find deposit at Llyn Cerrig Bach, Isle of Man, yielded wheel naves the ends of
which had been visibly overlapped and welded (Pleiner 1962, 194, fig. 36: 1, after Fox,
here Fig.26: 1 and 2. Admirable feats were archieved by blacksmiths during the Roman
period. The matter is of huge iron blocks or beams that have used in construction. The one
from Catterick Bridge (England, 2nd century AD) was welded together from at least 17
individual blooms (ca 7kg each, total weigt 250kg, see Tylecote 1986, 166, fig. 110; 1987,
230, fig. 7.5a). The beam from Corbridge (ancient Corstopitum, 3rd - 4th century AD)
was unfinished but demonstrated the method of gradual welding together some 15 blooms
in a vertical position in a former lime kiln which subsequently served as a welding furnace
(see Chapter VIII). It was about 1m high and weighed 156kg (Tylecote 1986, 156 - 157,
165 - 169, fig. 9; similar finds and their fragments are further listed in table 80; Tylecote
1987, 250, fig. 7.5b here Fig. 27). Another group of objects comes from the Roman
provinces Germania Superior and Germania Inferior. Baatz (1991) enumerates about ten
sites, mostly Roman military stations, equipped with baths: the beams served as supports
of the heating installations (caldaria, praefurnia). The piece from Kastell Saalburg was
investigated metallographically by Hauptmann and Maddin (o. c. 31sq.). The heaviest
block fragment weighed 140kg (the total weight of fragments found was 1.3 tons) and
consisted of several blooms of iron permeated with slag inclusions especially inside and
revealed coarse ferrite grains. In comparison with this evidence of such exhausting work
requiring the participation of two or more workers at the anvil, other welding operations
with wrought iron seem to be a mere playing.
(2) The welding of iron not only concerned the constructional operations of making
artefacts but also the preparation of semi-products and billets, i.e. the piling and faggoting
of flat bands and plates. This found application in the recycling processes of iron scrap.
Such a bundle of flat bands was easily peened. The individual plates could be of different
origin so that a heterogeneous composition of the metal has to be expected. Naturally,
it was hard to avoid the presence of slag inclusions in the welds; it depended on the
temperature and time devoted to the peening and on the tolerability in terms of loss on
the metal heated. Metallography of polished and etched samples of piled metal frequently
reveals laminated textures showing strips of different grain size, chains of slag and oxide
inclusions and alternating crystalline structures of wrought iron and steel. The piled
blocks were used in further shaping operations to produce artefacts. When well welded,
the lamination had a good effect on the quality of the forged piece as concerns its the
general strength. Bad welding might have been harmless, depending on the purpose of
the artefact. However, after ages of corrosion in soils the seams may be severely attacked
and the lamination appears on the excavated finds.
(3) The welding together of wrought iron and hard steel ranged among the masterful
techniques of hand forging in the past and was especially used in tools and weapon con-
struction, i.e. it aimed to equip the cutting-edges with hard material. The wrought iron
was heated to yellow or white glows whilst the carbon steel was in danger of decarbur-
ization an had to be heated merely to cherry red colours (400-500◦ C difference). The
60 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
metallography of the archaeological artefacts steadily brings the evidence of the perfect
mastering this difficult technique, at least on the part of the most experienced master
specialists. The application process of iron-to-steel welding demonstrates, in fact, the
high level that was achieved early on in the history of the smith’s craft (Fig. 26: 5 to
14) In Europe it was already known in the Hallstatt period (Early Iron Age) appearing,
however, rarely. An increase can be observed during the La Tène period but the applica-
tions were not yet standardized a occured in many varieties. A more uniform style began
to be outlined in the Roman period, particularly in the welding-on steel cutting-edges
(axes, chisels, knives). In the Middle Ages this technique dominated and determined
the production of quality blades up to the 15th century. It survived even longer in the
manufacture of axeheads.
(4) The acme of iron-to-steel welding is represented by striped damast and pattern-
welding. Both techniques profitted originally from the fact that apart of equipping the
cutting-edges with hard material the alternating of wrought iron and steel kinds results
in resistance against bending; i. e. they were rooted in the functional aspect. Gradually
they became display products: the use on naturally wet air caused spontaneous etching
effects - steel parts became darker, iron parts light grey, and phosporic band shone with a
silverish lustre. This alternating entered into intention in manufacture of knives, swords,
lanceheads an can be classified as stripe damast (Fig. 71: 4 - 4a).
Pattern-welding displayed different effects, as the term suggests, the light iron and
dark steel parts design of curved and undulated arches and S-lines created ornamental
impression. This was the effect of twisting iron and steel rods and wires subsequently
forged to a band. This was usually placed into the core part of the relevant blade, be
it sword, sax, or knife (Fig. 26: 7 - 8). In the case of knives the insertion of straight
phosphoric iron stripes multiplied the effect even after slight etching. The pattern was
indelible and became a mark of the quality and expense of the tool or weapon (so that
imitations performed with mere twisted iron elements appeared where genuine steel was
not available). See Chapter XI.
As mentioned above the cast damascene steel of the Orient represents entirely different
material for the making of blades (see Chapter XI) although light and dark tones appear
on them as well.

The kinds of welds might have been scarf, i.e. the ends were flatly placed one to the
other and pressed during hot work, or the one end was inserted into a split in the other
(V-joint). Butt welding in the manufacture of sword and knife blades to show the same
pattern on the reverse side as on the averse deserves particular attention.
The welding seams are visible on some archaeological finds on the surface but regularly
they are recognizable on metallographically investigated polished blocks. They may be
accompanied with slag inclusion chains - in dependance on work intensity. The boundaries
between wrought iron and hard steel may be more or less sharp. The longer welding, the
more carbon (pearlite) penetrates over the weld to the low carbon part. Experiments have
shown that after 10 minutes of intensive heating and forge welding all carbon travelled
over the seam (Pleiner 1973). In practice this was not the case in intentional conctruction
of the blade when the smith tried to work as quickly as possible; however, some pearlite
diffusion freqently occurs behind the weld in the wrought iron body. The weld seams are
FORGING OPERATIONS 61

Figure 24: Forging operations. G. Twisting: 1 Lovosice, Bohemia, Hallstatt period; 2


Kolı́n, Bohemia, La Tène period; 3 - 4 Novgorod, Russia, medieval; 5 Russia, site not
known, spiral auger, presumably medieval. H Splitting: 6 Rappenau, Germany, Hallstatt
period; 7 and 9 Stradonice oppidum, Bohemia; 8 Steinsburg oppidum, Germany; 10 Nov-
gorod, Russia, medieval, nail-puller; 11 Podhradı́ u Uh. Brodu, Moravia, early medieval,
arrowhead against cavalry; 12 Dobřichov-Pičhora, Bohemia, Romano-Barbarian. Various
authors and collections.
62 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 25: Forging operations. I Piercing: 1 Břasy, Bohemia, Hallstatt period, sword hilt;
2 Slovakia, site unknown, Scythian battle-axe; 3 and 8 Staré Hradisko oppidum, Bohemia;
4 and 9 Dobřichov-Pičhora, Bohemia, Romano-Barbarian; 5 Dobřichov- Třebická, dtto,
strike-a-fire; 6 Kappersberg, Germany, Roman; 7 Kúty, Slovakia, 9th century AD. Various
authors and museum collections.
FORGING OPERATIONS 63

Figure 26: Forging operations. J Hammer welding: 1 - 2 Llyn Cerrig Bach, Anglesey,
Early Iron Age period; 3 Staré Hradisko oppidum, Moravia; 4 Věteřov, Moravia, early
medieval; 5 and 6 Novgorod, Russia, early medieval (6 lyra-shaped strike-a-fire); 7 and 8
pattern-welded blades: 7 France, site not given, part of a sword blade; 8 Gdańsk, Poland,
early medieval; blade sections: 9 and 13, Novgorod, 10 and 14 Lutomiersk, Poland, early
medieval, 12 Lȩczyca, Poland, early medieval (dotted = steel). K Swage forging. L
Planishing of surface and sharpening of cutting-edge. Various authors and collections.
64 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
sometimes marked with so-called ‘white lines’ - in the heated and oxidizing surfaces a
segregation of certain elements took place: phosphorus, arsenic, nickel (Tylecote and
Thomsen 1973; Ustohal 2001; Hošek 2001). Most of the welds can be identified by com-
bination and evaluation of various traces and by using optical and analytical methods.
Their running through the objects produces fundamental data for interpretation of the
artefact construction.

Swaging and die forging

In contrast to free hand forging swaging represents an element of serial productiom. It


lends certain parts of iron artefacts a more or less uniform shape and size (rods, wire,
shanks, see Fig. 26: K. Virtual dies are rare finds among archaeological assemblages (see
Chapter VI) but many ancient anvils are equipped with grooves suitable for finishing
regular cross-section. As concerns the non-European weaponry a problem is presented by
the minute human heads fixed on the Luristan sword hilt pommels which must have been
shaped in a kind of matrix (stone?).

Planishing

Every ready made iron artefact shows surface unevenesses which have to be removed
by smoothing (Fig. 26: L and planishing under strokes of the hammer face or set ham-
mer. Nowadays the operation is carried out under lower temperatues (600-700 ◦ C), at
dark red or red heats. Finer planishing could be reached through the use of a file. Un-
fortunately, no planished surface has survived on the unearthed iron objects having been
substituted, in the case of cremation gifts, by a black patina and in general completely
destroyed by corrosion.

Sharpening and pointing

These operations are, in fact, kinds of drawing down cutting-edges, blades of knives,
swords, axeheads, chisels etc. or points (spear and lanceheads, nails, pokers, picks, see
Fig. 26: L). During pointing the workpiece could be steadily turned along the axis where
the symmetry and strihtness were wanted.

Shrinking-on

A different method of joining is the shrinking-on from higher temperatures when the
cooling metal is submitted to contraction. This method is not reserved only for modern
precize machinery. Every peasant cartwright when mounting iron tyres on wooden wheel
felloes or the cooper mounting iron rims on wooden barrels or buckets understood how to
achieve this by shrinking on. The Hallstatt period cartwrights nailed the tyre on wheel
rims as many princely graves with car remains attest (Bylany culture in Bohemia, HC,
gives good examples). However, the Celtic smiths in Britain and Gaul mastered the tech-
nique of shrinking-on iron tyres on wooden felloes and naves (Piggott 1983, 106, fig. 104;
215 - 216, figs 135 and 136). The tyres on the wheels of Classical Antiquity iconography
seem to be fastened without using nails but more detailed evidence is lacking. At all
events the shrinking-on of iron was already is rooted in cultures at least two and half
FORGING OPERATIONS 65
millennium ago.

Heat treatment

Some catgories of iron artefacts can be used after having been finished by shaping. Other
kinds have to be submitted to further processes, e.g. to those in which heating to tem-
peratures causing changes in the crystalline structure plays the main role. Therefore, the
processes are included in various methods of heat treatment. The aim was to meliorate
the metal properties according to the intended purpose.

Figure 27: Welding-together large iron blocks. The structural beams for Roman bath
heating installations in England. Left: Catterick Bridge, right: Corbridge-Corstopitum.
After Tylecote.
66 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Annealing

When is necessary to remove internal tensions in iron or steel, e.g after cold hammer-
ing, the brittle metal can be briefly annealed at the brown-red glow stage (600-650 ◦ C).
The recrystallization of grains returns the toughness to the metal. Annealing at higher
temperatures (800-950 ◦ C, according to the carbon content) produces increasing though-
ness. Steel implements which were overheated and those which are to be subsequently
quenched are annealed and moderately cooled; at about 721 ◦ C and after slow cooling
down to about 650 ◦ C the grains become finer, the lamellar pearlite is transformed to a
globular structure and the steel is softened.
Under severe heating conditions, e.g.during conflagration or when iron and steel object
are exposed to the heating during incineration burials, spontaneous annealing, overheat-
ing and decarburization depreciate the artefacts. When subsequently slowly cooled and
not forged the texture was transformed to coarse-grained. This was mostly the case in
the Romano-Barbarian Germanic culture milieu. The weapons, for example swords and
longer lanceheads were bent after annealing in the pyre to be put into the small burial pit.
It depended on actual place where the steel object was located during the fire; locally,
traces of hardening might have been preserved. Globular pearlite or a mixture of globular
and lamellar structures were observed on Hallstatt and Romano-Barbarian period forg-
ings (Rieth 1940, 155, fig. 93 - annealing in about 600 ◦ C after Lembeck); Pleiner 1962,
207).

Carburization

In contrast to annealing manufacturing practice demands that certain parts of tools and
weapons (surfaces, cutting-edges, points) should be harder. The more carbon in the steel,
the higher the hardness. It is well known that the bloomery process produced heteroge-
neously carburized and sometimes very hard metal. However, to have the harder parts
at disposal was not easy and practically uncontrollable. But the carbon content could be
introduced and/or increased secondarily - in critical working parts of tools. The body of
the instrument remains tough and its working edges get harder which warrant eminently
good properties.
The carbon content of steel with 0.02 - 0.15◦C could be increased by carburization or
cementation. At the yellow-red heat (about 900-950 ◦ C) the austenitic iron absorbs carbon
(through the dissolving of CO) from cementation substances, i.e. in principle charcoal
in the suiting place in the hearth or in special crucibles. Nitric substances like charred
leather, horn or hoof dust act as catalyzing elements. It is a long and time consuming
process - a 2mm thick carburized layer appears after some 12 hours. Thinner steeled
layers can be obtained in about 950 ◦ C when iron is stuck into the cementation mixture
after several hours. The carbon content (pearlite and pearlite-and-cementite structures)
decreases towards the inside of the artefact. The application in ancient and medieval
times has been attested without any doubt (see Pleiner 1962, 207 - 209; since that time
numbers of examples have been published; see Rehder 1981).
The most ancient evidence of the carburizing of iron, accidental or intentional, does not
come from Europe but from the Near East. Sets of knives, daggers and other implements
found in graves and settlements of the 11th - 10th centuries BC in Palestine (Mt. Adir,
Tell el-Far’ah, Tell Qiri, Taanach etc.) and Cyprus (Paleopahos-Skales, Idalion, Lapithos,
FORGING OPERATIONS 67
Amathus, Kition etc.) have been analysed. A survey of these finds investigated by
Maddin and Stech was presented by Waldbaum (1978, and explicitly 1999, 40 - 43, and
Appendix B, 48 - 50). European finds, especially those from the central territories of
the Continent date from the late Hallstatt period (6th - 5th centuries BC) through the
Romano-Barbarian and the early Medieval periods. A curved knife from princely grave
III of the Hallstatt period at Lovosice should be mentioned as an eloquent object forged
of pure iron and shallowly carburized from both sides of the cutting-edge (up to 0.4% C,
see Pleiner 1962,51 - 52, pl. X). Later, the most frequent cases of carburization may be
observed on iron artefacts not only from Romano-Barbarian sites in Bohemia (Sendražice,
Tišice etc.) and Poland (Bnin, Szwajcaria - up to 0.75% C, Tarnówek - all of the surface
carburized, see Pleiner, o.c. 123, quoting contributions by Piaskowski as to Polish sites)
but also from the Roman provinces in Britain, ibidem.
The cutting-edge of a blade was stuck into the caburizing milieu both vertically or
obliquely as the distribution of pearlite-containig areas shows; the back and the body
of the blade used to be protected against carburization e.g. by clay. The all-surface
carburization like at Tarnówek could have bee achieved either by sinking all the blade into
the charcoal-filled hearth or placing in special crucibles filled with charcoal an heating it
in a hearth or furnace. This, relatively rare, method was applied e.g in early Medieval
Russia, sites of Ryazan’, Glazov, Podbolot’ye, see Kolchin 1953, 74, fig. 35: 18 and 19;
110, fig. 72: 8). Some of the crucible finds did not serve for the smelting of iron ore as was
previously supposed but for the thorough carburizing of small iron sheets or plates. In
that way, the steel needed would be produced more economically and used in welding-on
techniques in the tool construction. In the East the carburizing in fuel-filled crucibles
was the way to the production of cast steel: the cementation was so rasant that the iron
particles got molten and were formed into ingots of the wootz type of steel.
With the intensification of ironworking in advanced civilizations the use of this time
consuming technique receded and in the High Middle Ages was not practically applied in
tool making (the handbook of the 11th century, called Theophilus, describes the carbu
-rization of files using fat and leather as catalyzators). However, the cementation process
has not been forgotten in modern industry where case-hardening furnaces produce e.g.
carburized tooth gears.

Quench-hardening

The hardness of carbon steel containing more than 0.35 - 0.4% C increases up to three
or four times when rapidly cooled (Fig. 28). Under conditions of an abrupt fall of tem-
perature the carbon cannot be eliminated from the austenitic stage like cementite in the
pearlitic structure. It crystallizes in form of lens-shaped particles which appear in section
as needles - this is martensite, an instable solid solution of carbon in α-iron, very hard
and very brittle. The temperature from which the steel is being quenched depends on the
carbon content: about 900 ◦ C (orange glow) for steels containing ca 0.4% C, about 830◦ C
(bright cherry) for those with ca 0.5%, 815 ◦ C for 0.6% and 780-770 ◦ C for hypereutectoid
steels (over 0.8% C).
The glowing piece of steel has to be dipped into a cooling agent: cold water, may be
mixed with an acid (not suitable for steels above 0.9% C). Salty, hot or soapy water has
a more mild effect than oily water or oil and a stream of cool air. Practice shows that
water that has once been used in the quenching process provides a more uniform hardness.
68 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 28: Comparison of hardness of processed iron-carbon and copper-tin alloys. After
C.S. Smith.

Because of that the masters did not did not pour it out and they even watched jealously
over it (Weber 1946, 78; Hlubuček 1953, 3 - 4). It is interesting that the Czech and Russian
terms for quenching kalenı́ and zakalka involve the stem kal, mud, so working with muddy
water. An all-steel artefact, grasped in tongs, used to be vertically dipped position and
moved up and down to expel the bubbles adhering to the surface of the quenched object;
the bubbles might have caused irregularties. They were thoroughly hardened (e.g. the
hammer from the Celtic oppidum of Stradonice, see Pleiner 1962, 83, pl. XIX). Other
tools, eg. knives, axeheads and the like were partially quenched only along the cutting
line. When taken out of the coolant the absorbed heat spread from the tool body over the
edge where tempered and unquenched structures appeared allowing the gradual transition
from the very hard to tough zones. In respect of the working properties it was an ideal
distribution of hardness and toughness and the metallographically observed structures
attest it very frequently on objects from all of the chronological periods and cultural
provinces. Good examples are offered by Early Iron Age or La Tène period axeheads e.g.
from Ireland (Toome Bar and Glenariff, Co. Antrim, see Scott 1990, 52 - 57, pl. 3.3.3
and 3.3.4) or from the Manching oppidum (No 1963/257, see Pleiner, sine anno, sample
715. Medieval items should be consulted in Pleiner 1962, 210, with references also to
FORGING OPERATIONS 69
Schürmann and Schroer, and Piaskowski).
When the cooling is slower, at temperatures lower than necessary for forming marten-
site (down to about 500 ◦ C) another phase results in austenite disintegration - the bainite,
a dispersed precipitation of fine ferrite needles and cementite. Below the temperature
mentioned it is a hard and stable structure and not brittle; it has been observed in many
heat-treated archaeological artefacts.
The quenching of tools and weapons in moderately acting liquids like oil or boiling
water decreases the brittleness, especially in case of hard or even eutectoid steels (exam-
ples see Pleiner, ibidem, notes 59 to 64; the matter is of knives, axeheads, chisels, swords,
sabres form the Iron Age up to Middle Ages). Apart remaining martensite appear struc-
tures from the family of fine pearlite (e.g.troostite and sorbite appear).

Tempering

The additional heating of quenched steel objects to 200-450 ◦ C (e.g. axeheads and in-
struments exposed to heavy shocks) causes structural changes in which the martensite is
transformed to fine pearlitic phases known by the infrequently described terms of tem-
pered troostite (dark fine pearlitic spots in the martensite matrix) and sorbite (a mixture
of pearlite and cementite in the form of fine globules or needles). The hardness slightly
decreases but not drastically while the higher toughness is favourable during the use. In
small smithies the tempering was performed on smoothed parts of tools put on an iron
plate and heated to 200-300 ◦ C (or more) according to the required property of the steel.
The temperature can be estimated on the base of colours if tiny oxide layers developed
on the tool surface:

C colour

210 whitish yellow


220 - 240 bright yellow, yellow (chisels, knives, planes)
240 - 260 dark yellow to brownish red (hammers, augers etc. )
270 - 280 purple to violet (drifts, smith’s chisels)
290 - 320 blue to grey (rivet irons, axeheads, scythes)

Table 2: Heating colours of tempered steel.

An intensive tempering could be performed even at higher temperatures (500-600 ◦ C).


After heating the object has to be cooled rapidly again to avoid further annealing pro-
cesses.
Traces of this technique appeared on steels of the Roman period and it seems that
after AD 500 it became common. According to Kolchin (1953, 178) the early Russian
axeheads were tempered at relatively high temperatures (500-600 ◦ C) but scythes and
sickles at a lower heating region of about 300 ◦ C) sofar the troostitic and sorbitic phases
were classified. Anyway, the distinguishing of these structures from those having been
developed in mild coolants may present difficulties.
Both quenching and tempering required much experience and skill from ancient smiths.
70 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
The problem of nitridation

Under high temperatures nitrogen enters into reaction with iron and produces hard ni-
trides which could be dissolved in solid solution of iron. When oversaturated, the nitrides
precipitate as iron nitrides. A lot of metallographically examined iron taken from incin-
eration burials show dark shorter and longer needles of nitrides (γ’-Fe4 N3 , α’-Fe16 N2 ).
Piaskowski (1959b, following Booker, Norbury and Sutton) introduced them into ar-
chaometallurgical literature. The particles are, however, unstable - they disapear when
iron is reheated and more rapidly cooled. These nitrides in ferritic structures do not rep-
resent any nitridation. Nitrogen penetrates into iron during the smelt from an air-blast
(air contains, apart some inert gases, 78% N and 21% O). It did not influence the hardness
of the metal (Tylecote 1986, 193; Scott 1990, 16, 19).
Intentional nitridation as a method of hardening steel is practiced in modern metallur-
gical industry in special installations (nitridation furnaces) at temperatures 500-600 ◦ C
in an ammonia medium during tens of hours. Speculations about the use of intentional
nitridation in the past are rooted in an episode in Germanic mythology - a place in the
Amelungenlied where a renowned master, Weyland the Smith, manufactured the super-
sword Mimung. He took King Neiding’s sword and pulverized it with file. He mixed
the filings with milk and flower to dough and fed with it the poultry (Mastvögel). Five
days afterwards he smelted the droppings and dung and on the seventh day he created
Mimung, the ‘king among swords’. The application of nitric dung brought about the idea
of nitridation. In fact, nitrogeneous substances catalyzed the absorption of carbon in the
steel, and phosphorus as well - both were favourable for the increasing of hardness. No
archaeometallurgical evidence of steel nitridation is available. Less skeptical seems to be
the opinion by Willen et al. (1976).

Except of the last discussed theme about nitridation all of the forming and heat-treating
technologies were known by traditional European blacksmiths and smiths and represent
a part of the metallurgical tradition and heritage.
Chapter VI

THE SMITH’S TOOLS

Effective tools were crucial devices for those who worked, hot or cold, hard materials
like metals: copper and copper alloys, or iron and steel. Smiths who began the working
of iron disposed with technological preconditions: founders and bronzeworkers mastered
forging processes shaping non-ferrous metals with bronze tools, especially during the late
Bronze Age and early Hallstatt period. Bronze anvils, very often equipped with horns for
bending rods and metal bands, bronze hammers (mostly socketed), and rarely bronze files,
belonged to the equipment of their workshops in France, Germany, Switzerland, Austria
and central Europe (Ohlhaver 1939, 103 - 1ll, Pl. 1 to 6; Mohen 1980).

The find complexes

Omitting the working of iron with stone on stone (some examples of which were still
observed in 1950’s in Africa, see Coghlan 1956, 109 - 110; Gardi 1954), we turn the
attention to iron smithing tools as registered among the archaeological evidence. However,
stone anvils quite frequently appear on the scene (see Chapters VIII and IX).
The identification of smiths’ iron tools among archaeological finds may present some
difficulties. Some heavier items allude to the work of the blacksmith proper (in German
‘Grobschmied’ while other sets of e.g. hammers, tongs and anvils together with files, nail
irons and drawing dies belonged to the equipment of smiths producing smaller objects
and who were engaged in shaping non-ferrous metals as well, in German ‘Feinschmied’).
Then, in the course of a smith’s specialization, some specific tools played a crucial role:
files in the work of the locksmith, nail irons in the production of nailmakers, wire broachs
for drawing non-ferrous and later, in the Middle Ages, iron wire.
Nonetheless, tools like hammers were not necessarily smithing tools, they might served
in any the operations involving strokes. Some light onto their identification could be pro-
vided by considering the finds’ circumstances and the milieu in which they appeared.

Tools from smithies

When abandoned in peace-time, smithies yielded a few examples of tools found during
their excavation. It is not possible to deal with all of the finds on these pages. Neverthe-
less, some cases indeed deserve attention.
At least in fourteen smithies dating from the prehistoric up to medieval times smithing
tools or their fragments were found (as shown in Chapter IX). Here, two cases should be
mentioned. First, the mysterious case of the HD period (5th century BC) at the Býčı́
skála cave in the Moravian Karst. In the so-called smithy or symbolical smithy or deposit
there three heavy sledge hammers were found by Wankel (1882, see also Ohlhaver 1939,
115; Pleiner 1962, 62, Fig. 20), and, in addition, two small block anvils and a half of
a bipyramidal bar or ingot, and pincer tongs. The sledge hammers whith shaft holes
bear traces of use but their size and weight (6 - 8kg) make a strong impression; the date,
relevant culture and the presence of a mass burial or sacrifice in the close vicinity in the
same cave allude to extraordiry circumstances.
72 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
The second example is the smithy in the pentagonal tower of the fort of Sadovec, Bul-
garia, from the 2nd half of the 6th century AD. There, a large set of smithing tools came
to light: a horned and a block anvil, three hammers, six small tongs, part a pair of pincer
tongs, a chisel and two sheet-shears were left, possibly due to dramatic circumstances of
evacuation (Welkow 1935; Ohlhaver 1939, 125 - 126, Fig. 52).

Smithing tools from settlements

Ironworking shops produced iron artefacts in villages, oppida, military camps, medieval
towns and rural settlements which is attested by smithing slags and other wastes (see
Chapter VIII). The same has to be applied to stray finds in the layers of these localities
which came from abandoned, destroyed and not recognized smithies. The number of these
smithing tools is enormous and there is no possibility to discuss them in detail within this
context.
Larger assemblages come from the La Tène period (mostly fortified) sites like Manch-
ing (Jacobi 1974, 5 - 27), Bern-Tiefenau (Tschumi 1953), Sanzeno (Nothdurfer 1979, 35
- 38), Sarmizegetusa (Glodariu 1980) etc. Then, from the Roman towns and legionary,
e.g. Magdalensberg in Austria (Dolenz 1998, 159 - 177), Cáceres el Viejo in Spain (Ulbert
1984,158 - 1560, Augsburg-Oberhausen in Germany (Hübener 1982), 44 - 45), etc., and
early medieval towns like Staraya Ladoga and Novgorod Velikiy in N Russia (Kirpichnikov
et. al 1981; Kolchin 1959, 58 sq., respectively). More sites will be mentioned below when
discussing individual categories of tools.

Smithing tools from graves

A specific category are smithing or blacksmiths’ tools in graves. Their presence among the
burial gifts caused discussion about the social position of the buried men. This changed
during time. In an historical context smiths worked for the court (metalworking was
controlled by the upper class especially at the beginning and during the early periods of
the Bronze and Iron Age), or as slaves, or as masters who employed other smiths (an-
cient Greece). They worked as military smiths (Rome) or as members of domestic estates
(villae rusticae, Germanic farms). Franchised libertini and to certain extent free smiths
lived and worked in the Roman world. They were not buried with their implements but
the principal emblems of their profession figure on their tombstones: anvils, hammers
and tongs (for illustrations see above all Esperandieu 1910 - 1938). In the Middle Ages
free smiths worked in towns. On the contrary, in the countryside the feudal relations
prevailed. In addition: itinerant smiths (incl. Gypsies) existed until the 20th century.
Graves containing ironworking tools number many hundred, partly in the La Tène
period in Central Europe and especially in Romano-Barbarian and Viking regions in the
North. (Ohlhaver 1939; Blindheim 1963; Malinowski 1952; Kokowski 1981; Henning 1991;
Müller-Wille 1983):
The so-called ‘smith’s graves’ should be characterized in three subcategories. These
of the first group contained some smithing implements, possibly accompanied with some
personal things, individual metal objects, ornaments, pottery vessels. The buried masters
were equipped, after their decease, with symbols of their profession; this attests to their
significance within the relevant community, although their actual social position remains
veiled. It is appropriate to quote some examples. La Tène period: St. Georgen in
THE SMITH’S TOOLS 73
Austria, tongs, a hammer, a file, shears, a fibula (Taus 1963); Leipzig-Thekla in central
Germany, tongs, a chisel, shears, girdle gear, a vessel (Moschkau 1962); Boddin grave
4, Germany, an anvil, girdle gear, a female burial? (Keiling 1972 for other references
see Henning 1991, 66 - 70). Dacian Porolissum, Romania, an anvil (Macrea and Rusu
1966). Romano-Barbarian period and civilization (mostly incineration burials): in Poland
Rza̧dz, a hammer, 5 files, 2 knives; Kalisz, tongs, 2 hammers, 3 files, an anvil (Malinowski
1953, figs. 3 and 5); Zadowice, tongs, hammer, chisel, file, iron wire (Kaszewska 1981,
fig. 2); Szaniec, tongs, an anvil, a pottery vessel (Skurczyński 1958), In Germany Pleets,
a hand hammer (Keiling 1987) and Hüchelhoven, tongs, 2 hammers, a nail, unidentified
implements, disturbed (Weinand and Piepers 1967).
Later, in the early Middle Ages, a set of blacksmith’s tools was discovered in a grave
from Skredtveit, Norway (2 tongs, 3 set hammers, 2 hammers, 2 chisels, wire drawing die,
nail iron, an iron bar (Müller-Wille 1983, 257, fig. 22). Grave gifts of that kind are hardly
to be expected in medieval graves in countries where strict Christianity ruled. However,
two examples shall be presented: the inhumation grave at the Danish Lejre (island of
Seeland) dating from the Viking period (a hammer and a file were deposited at the legs
of the skeleton and tongs at his head (ibidem, 150, fig 7: 3 - 4). From Redikar (Russia,
9th - 10th centuries AD) Lunegov (1956, fig. 41: 5 - 6) presents a grave with tongs and
smith’s hammer.
The problem of the social status of the smith becames more acute when the second sub-
category is taken into account - the smith’ s graves containing weapons. The buried men
need not have belonged to the caste of warriors but they were presumably free members
of the comunity; Henning (1991) would count them as belonging to the armed folk. In
some graves modest arrowheads were found, but in others also lances, spears, and saxes,
shield bosses, rarely swords, helmets and horse-gear. The tumulus from Celles (Auvergne,
France, 1st century BC) contained a hammer, files and rasps, a draw-knife and a lance-
head. Ohlhaver (1939, 195) throws doubt on the function of the above implements saying
that they could be held to have been the property of a woodworking craftsman. Other-
wise the majority of the ‘smith’s graves’ with weapons comes from regions inhabited by
Germanic tribes and date from the Romano-Barbarian and Migration periods. Examples
(mostly incinerations): Dessau, central Germany, 2 tongs, a file, a lancehead, a pottery
vessel (Ohlhaver 1939, 123, Pl. 12: 1). In Poland Wlostowice-Pulawy grave 15, a ham-
mer, a file, a lancehead, 3 arrowheads, a shield boss, a girdle buckle (Kokowski 1981,, 202,
fig. 8), Korytnica, tongs, a hammer, a file, 2 knives, a girdle buckle, a fragment of a pair
of shears, an arrowhead (Malinowski 1951/1952, fig. 2). Among the incineration graves
of the cemetery at Wesólki - three graves (Nos 3, 36 and 45) contained smithing tools
from which grave 36 was equipped - apart from tongs, hammer and anvil, 3 files, with a
lancehead and a shield boss (Da̧browscy 1967, fig. 43, Pl. IV: 5). A specific case is repre-
sented by the inhumation from Gannor, Gotland, Sweden, where a sword and shield boss
were buried at the left side of the skeleton, whilst tongs and a hammer were deposited,
perhaps originally wrapped, at the feet (Müller-Wille 1983, fig. 7: 1 - 2). Two Migration
period complexed deserve attention. Grave 6 at Poysdorf, Lower Austria, yielded 2 tongs,
2 hammers an anvil, stone casting moulds, and a shield boss and sax. The presumably
Gepidic inhumation grave X at Mezöbánd in Romania contained 2 tongs, 2 hammers,
wire/nail iron, bronze and iron mountings, and a helmet, not to mention various metal
strips, girdle buckles etc. (Ohlhaver 1939, 123 - 124, 127, Pl. 13: 1, Pl. 14).
In certain early medieval graves of northern Scandinavia not only smithing tools and
74 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
weapons appeared but also other implements usable in woodworking (rasp, axes, augers)
and even sickles as parts of agricultural equippment. Let us mention Haglebuvand in Nor-
way with tongs, a helmet, an arrowhead and a bridle bit (Ohlhaver 1939, 143; Müller-Wille
1983, 252, fig 16: 16 - 35) or Transtrand in the Swedish Dalarna with tongs, 3 hammers, a
file, an axe, a sword, a spear (Serning 1965, 76, fig. 10). The largest assembly comes from
Bygland, in Norway. Not only blacksmith’s tools were identified in the grave inventory (2
tongs, 4 set hammers, 1 handhammer, 2 files, nail iron and wire-drawing die, sheet shears,
chisels, forge spoon) but also stone casting mould, keys, padlocks, rasps, scythes, sickles
and weapons: lanceheads (Blindheim 1963, 25 - 80; Müller-Wille 1983, 257 - 259, fig. 23).
Blindheim expressed the opinion that the buried craftsman was equipped with his tools
and with his products as well. Another idea emerges: the buried personality could be a
landlord whose grave gifts signalize the activity of his artisans and farmers.

Smithing tools from hoards

Hoards of iron objects rarely contain blacksmith’s or smithing tools as a specific cate-
gory of find. Mostly they were a part of an inventory comprising other artefacts as well.
In certain cases, the smithing tools represent a minority or even single pieces among tens
or hundreds objects. The problem is what was the reason to hide these things and elimi-
nate them from the circulation. Was it the property of a craftsman or a votive deposit?
(see Chapter XII). Some of the large hoards including smithing tools may be considered as
a cultural heritage, some other assemblages were discussed during the recent presentation
of the field work.
Several examples should, however, be mentioned here. Again, apart of the mysterious
set of sledge hammers and block anvils found at the Býčı́ skála cave (see above) attention
has to be paid to the La Tène period hoard from Golling-Nikolausburg in Austria where
blacksmith’s tools (2 tongs, a hammer, an anvil, a forge spoon) were accompanied by an
iron band (Moosleitner and Urbanek 1991) and to the hoard from Kappel, Württenberg,
Germany, of the same period (tongs, set hammers, large fire-dogs, a tripod, a scythe,
bronze vessels, see Fischer 1959). A relatively well-known assemblage is that found in
a grovel pit at Bern-Tiefenau in Switzerland (3 tongs, a drift, a hammer, 2 horseshoes,
see recently F. Müller 1990). In a La Tène period hoard from Bezdědovice, S Bohemia,
were found 2 hammers and a forge spoon among numerous metal objects, tools and
mountings (Michálek 1999). Waltham Abbey, Essex, dominates the British sites related
to the Roman period (block anvils and a hammer with forming die grooves, six tongs, see
Manning 1971 who holds it for a ritual deposit). Another hoard in Britain, Bulberry in
Dorset, yielded a hammer, a fire-dog, an iron anchor and a bronze vessel, all presumably
hidden in a wooden chest (Cunliffe 1972, fig. 6: 16). The Roman hoard from Sandy,
Bedfordshire, contained a block anvil an 2 field anvils from which one was interpreted as
a cobbler’s last. In the frontier post at Newstead, Scotland, a hoard was unearthed with
5 hammers, 2 field anvils, 2 tongs etc. (Ohlhaver 1939, 122 - 123,Pl. 11, after Curle).
Eastern Europe: Dombóvár is a Roman site in Hungary where a hoard of 87 items included
tongs, a hammer, nail-iron and a pair of compasses (Gaál 1982, fig. 3). Ošaniči (Daors
in Herzegovina, 2nd century BC), Lozna and Marculani in Romania represent important
hoards. At the Illyrian Ošaniči a large set of about 17 smithing tools (anvils, hammers,
tongs, files) was found among 245 iron artefacts (Marić 1979). Block anvils, hammers,
vice-clamps, a forge spoon were recognized within an assemblage of 56 items from Lozna
THE SMITH’S TOOLS 75
(Dacian milieu), see Teodor (1971). At Marculani an iron anvil was found in a hoard
numbering 127 items (Glodariu et. al. 1970). In the Roman province Germania Superior
(SW Germany and E France) should be mentioned Selz, ancient Saletio, in Alsace, Gallo-
Roman (2nd - 4th centuries AD), with hammers, a horned anvil, tongs, a chisel and file
and with woodworking implements and soldering devices Schaeffer 1927) and Heidenberg
bei Kreimbach (Bavaria, 3rd - 4th centuries AD) - horned and block anvils, tongs with
differing jaws, a forge spoon (Ohlhaver 1939, 117 - 118; Gaitzsch 1980, 37, Fig. 9);
As concerns the non-Roman, barbarian lands, the hoard from Tluste, Poland, contained
blacksmith’s tols (an anvil, tongs, a hammer) and agricultural implements (Waluś 1979).
And the Danish site of Vimose (Fn̈en) should not be omitted as this represents not any
hoard as it represents not merely any hoard, but a large sacrified moor deposit with
weapons - Ohlhaver (1939, 128 - 129) describes, after Engelhardt, tongs and hammers
which were also there during the 19th century (see Chapter 10).
The Migration period is well represented by the site of Runder Berg bei Urach, Württenberg.
Various kinds of hammers and tongs were among the agricultural implements and chains
discovered there (Koch 1988.
Two hoards dated to the Viking Middle Ages cannot be left out of our examples. There
were found in Tjele in Jutland in 1850 and contained a set of tongs, an anvil, hammers, 3
nail or wire drawing irons, sheet shears, a riveting iron, a die, 5 files together with axes,
keys, sickles, and a casting mould (Ohlhaver 1939, 130, Pl. 16; Munksgaard 1982). From
early medieval times Halleby (Denmark, Seeland) deserves a mention. Here a hoard with
tongs, anvil, sheet-shears was uncovered (Ohlhaver 1939, 128). All of the Scandinavian
hoards are crowned by the implement chest of Mästermyr on Gotland in Sweden (Müller-
Wille 1977; Thålin-Bergman 1979, 101 - 115; eadem 1983); The remains of a wooden
chest containing a number of iron objects was found in the southern part of the island in
marshes of that name. It contained above all smith’s and carpenter’s tools, not to mention
balances, two iron bars and a piece of lead. In total, 80 metal artefacts were hidden in
the moor either for ritual reasons or when a master craftsman met with an accident while
on his travels.
Eastern Europe should be represented by a medieval hoard of weapons and blacksmith’s
tools from Dragoslaveni, Romania, which comprised 6 tongs, 2 hammers, an anvil and
metal sheet-shears (Comşa and Constantinescu 1969, 427, fig. 1).

The tools

The tools destined to perform smithing operations were used, in the case of iron, mainly
for hot work and to a smaller extent for the cold working (cold hammering) - specific
devices were intended as cold working tools for fine and finishing operations (files, wire
drawing irons, metal sheet-shears).
The basic tools to be considered are hammers, tongs, and anvils. Nevertheless, one has
to take into account that shaft hammers have to be pierced to make a shaft hole, and this
requires a chisel or punch. Hinged tongs must be also pierced with a drift and, moreover,
it was necessary to connect both arms with a rivet - with the aid of other specific tools.
Therefore, hammers, tongs, anvils, chisels, drifts and rivet irons belong to categories of
tools which have been in use at least since the 6th century BC in Greece as it is attested
on painted vases (Fig. 29).
76 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 29: Smiths and their tools on Greek painted vases. 1 London lekythos, scene of a
forge; 2 bowl in the Museum at Mannheim, travelling Haphaestus; 3 dish, Antikmuseum
at Berlin, sitting smith at an anvil; 4 amphora in the Boston Museum of Fine Arts,
smithing scene. 1 and 4 black-figured vases, end of the 6th century BC, 2 and red-figured
vases, 5th century BC. 1 Pleiner with references, 2 and 3 after Brommer, 4 after Ziomecki.

A developed tool set includes varieties of hammers, tongs (universal, with shaped jaws)
and anvils (block, horned, those equipped with nail holes and swage grooves), and various
types of instruments serving to produce nails separately and to draw - non-ferrous - wire,
punches and stampas, and files for abrasive work.

Hammers

Hammers were and are tools which could and can shape iron (mostly hot, in plastic
austenitic state over 900 ◦ C) or cold, causing deformed grains, not rarely with traces of
slip surfaces known as Neumann bands.

Hand hammers

A universal means of applying strokes to various materials, and also shaping metals
THE SMITH’S TOOLS 77
through strokes; they weigh about 0.5kg to 2kg. The heavier the hammer, the shorter be
the path of the impact can be. Then, the work is slower. Hand hammers are operated
by one arm of a man. Hammer heads are positioned in the right angle to the (usually)
wooden shaft set into a rectangular, oval or circular hole. One side of the tool is thinner
and is called pene or peene (in Germann Finne) serving in drawing out processes. The
opposite flat working side is face (in German Bahn) for flattening and planishing the metal
surface.
Iron hammers of this kind (not to mention socketed Bronze Age bronze hammers used
in the non-ferrous metalworking) have been in use since the antiquity and their shape, in
principle, did not changed until now.
The earliest iron artefacts reflect rather fine work although the factual evidence of the
use of hand hammers has been, up to date, relatively late (Classical period in Greece,
La Tène period in central and western Europe). A hammer head, 13.5 cm long, comes
from Olynthus in Greece (4th century BC, and another one (11.5cm) from the legionary
camp at Cáceres el Viejo in Spain (Republican, 2th century BC). Then, from the La Tène
period onwards, the finds of hand hammer heads are innumerable. These implements will
be commented on and some examples will be presented.
A remark has to be made about the shaft holes of ancient hammer heads. They were
pierced, when hot, with a chisel which produced a rectangular hole, or with a round-
sectioned drift leaving a circular or oval shaft hole. The above mentioned two hammer
heads were round-pierced but the sides were smoothly levelled and planished. This was
not the case with other, especially Roman hammer heads with circular holes: the sides
remained swelled, in many cases with swells intentionally slightly edged. This ‘Roman’
style was practiced in some regions up to the Middle Ages (Fig. 30). In the literature,
dimensions are usually only given when describing tools of (10cm to 20cm, rarely even
more), weights only exceptionally appear (Magdalensberg, 0.2 kg to 1.2kg, see Dolenz
1998, 160 - 161).
Hand hammers with planished sides around the (regularly rectangular) shaft hole are
principally encountered in northern parts of Europe (Fig. 31). They measure about 10cm
to 15cm in length. The shafts were wooden and in some shaft holes traces of this material
remained. Hammers equipped with iron shaft are rare. In the princely graves at Vendel,
Sweden (7th century AD) several such implements were found (Ohlhaver 1939, 135, fig.
54, 25cm long; Pl. 18: 2 - 6, up to 31cm, see Fig. 32). Normally, the length of wooden
handles must have been 40cm to 50cm. A medieval hammer from Bargun in Germany
deserves particular attention. Its head of pierced in a unique way: the shaft hole diverges
into two holes on the upper surface (Fig. 31: 9, see Schuldt 1980).
When the peene runs parallel with the shaft (like an axe), the implement is called
cross-peene hammer. A nice example represents an item from the Roman hoard from
Kreimbach, Germany (Fig. 30: 6).
The shaftless hammer documented from Africa by Gardi (1954) has to be seen as a
pecularity. An about 30cm long iron bar was applied in the manufacture of sickles. It
would be hopeless to look for identifiable shaftless hammers among ancient and prehistoric
finds.
Hammers were exposed to hard shocks and the working surface had to be steeled. A
small metalworking hammer from the La Tène oppidum at Stradonice, Bohemia, was
made of steel and revealed martensitic structures (Pleiner 1962, 83, Pl. XIX). Fell (1993)
investigated 4 Iron Age hammer heads from Bredon Hill in England and discovered he-
78 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 30: Selection of Roman and Roman-style hammers. 1 Sanzeno, N Italy, 1st century
BC, length 16 cm; 2 - 3 Cáceres el Viejo, Spain, legionary camp, both about 10cm; 4 - 5
Magdalensberg, Austria, early Roman, 11cm and 24 cm respectively; 6 Kreimbach, SW
Germany, Roman, a cross-peene hammer, 25.5cm; 7 Hüchelshoven, Germany, 13 cm; 8 -
9 Moosberg, Germany, 18cm with double peene, and 20cm; 10 Dragoslaveni, Romania,
medieval, 13cm; 11 Kolyu, Bulgaria, 36cm. 1 after Nothdurfer, 2 - 3 after Ulbert, 4 - 5
after Dolenz, 6 after Ohlhaver, 7 after Weinand et al., 8 - 9 after Garbsch, 10 after Comşa,
11 after Changova.
THE SMITH’S TOOLS 79

Figure 31: Hand hammers (examples): 1 Kappel, SW Germany, from a La Tène period
hoard; 8 Magdalensberg, Austria, Roman, 2 Siemianice, Poland, from an incineration
grave, Romano-Barbarian; 3 Wesólki, Poland, Romano-Barbarian incineration grave 36;
Viking period: 4 Bredsundsnäsle, Sweden, from grave 3; Viking period: 5 Tjele, Jutland,
from a hoard; 6 Bryn, Norway, from a burial mound; Slavic and early medieval finds:
7 Vladimir, Russia, from a burial mound; 9 Bargun, Germany, 12th century AD; 10
Novgorod, Russia, 13th century. 1 after Fischer, 2 and 3 after Malinowski, 4 after Serning,
5 and 6 after Müller-Wille, 7, and 10 after Kolchin, 8 after Dolenz, 9 after Schuldt.
80 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 32: Hammers equipped with iron shafts. 1 Xanten, ancient Vetera, NW Germany,
Roman (nail-puller, 19cm); 2 Vendel, Sweden, early medieval, 31cm; 3 Aska, Sweden,
early medieval, 18cm. 1 after Gaitzsch, 2 and 3 after Ohlhaver.
THE SMITH’S TOOLS 81
terogeneously carburized heat treated steel (martensite, bainite, ferrite; hammer 2 was
of quench-hardened steel, hammer 1 was air-cooled - pearlite, ferritic network in Wid-
manstätten arrangement).

Sledge hammers

Sledge hammers weighing 3kg - 6kg or even more were operated with both hands and
required an assistant to the forge master when working heavier hot pieces of iron. The
action was already depicted on Greek painted vases (late 6th century BC, see Fig. 29:
1 and 4). In the archaeological evidence sledge hammers appeared relatively early. Four
heavy sledge hammers with rectangular or oval shaft hole were found within the HD find
complex (5th century BC) from the Býčı́ skála cave in Moravia; the objects were in differ-
ent stages of preservation (one was broken, another one has been clearly used, the weight
more than 6kg, length 13cm - 16cm, faces between 5cm x 5cm and 7cm x 7cm, one of the
pieces see Fig. 33: 1. Another heavy hammer, nearly of the same period (Late Lusatian
culture) was found in the hillfort of Wicina in S Poland (Bukowski 1981a, 71 - 72, fig. 4).
Various sledge hammers are presented, as examples, from the Iron Age up to the Middle
Ages in Fig. 33.

Set hammers

Work with a set hammer requires the aid of an assistant as well. Set hammers were
applied in planishing the grooved and uneven surface of metal which had been drawn out,
prolonged or widened, or in the shouldering of the hot material. The shaft hole of the set
hammer is located excentrically close to the face which receives the strokes of the sledge
hammer; the working part is, contrary to the peene, flat. Set hammers were definitely
used since the La Tène period and are 10cm to 13cm long (some models see Fig. 34).
Curiously enough, some ancient Greek and Roman iconographical sources show the use
of excentrically positioned shaft on hammers which resemble set tools but were operated
as hand hammers (Fig. 29: 3; Fig. 36).

Tongs

A smith shaping shorter or smaller iron pieces in hot work needs an instrument which
is able to grip or clasp the glowing metal. Ethnologists reported that some African iron
smiths worked with green whit tongs used as tweezers (Coghlan 1956, 126, after Jeffreys).
The whit tongs are predecessors of pincer tongs made of metal.

Pincer tongs

These implements were not only smith’s device but were used by metalworking founders
as well because they were forced to hold and manipulate hot crucibles. They helped
themselves with green whit tweezers or with (bronze) pincer tongs which originated in
the Near East and Eastern Mediterranean at least during the 2nd millennium BC (Jock-
enhövel 2001, 92 - 93). Later, they were often applied in smithing; a hand hammer and
pincer tongs were, in some cases, attributes of Hephaestus (a bronze stauette in Berlin
collection, see Saska and Groh 1949, 45, fig. 17, here Fig. 36). The fund of European
82 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 33: Sledge hammers. 1 Býčı́ skála cave, late Hallstatt period; 2 Krivodol, Bulgaria,
Hallstatt period, from a hoard; 3 Serbia, site unknown, Roman; 4 Selz, E France, Roman,
from a hoard; 6 Sarmizegetusa, Romania, early Dacian; early medieval sites: 5 Ekimaucy
hillfort, Russia; 7 Preslav, Bulgaria. 1 after Pleiner, 2 after Nikolov, 3 after Popović, 4
after Ohlhaver, 5 after Fedorov, 6 after Glodariu, 7 after Chengova.
THE SMITH’S TOOLS 83

Figure 34: Early set hammers. La Tène period: 1 Kappel, SW Germany, from a hoard; 2
Golling-Nikolausburg, from a hoard; 3 Dürrnberg near Hallein, Austria; 4 Lozna, Roma-
nia, about BC/AD; 5 Manching oppidum, Bavaria; 6 Skredtveit, Norway, Early Medieval;
7 Moosberg, S Germany, Roman; 8 Brandsunsnäset, Dalarna in Sweden, Viking perod,
from grave 3; 9 Bygland, Norway, Viking period, from a hoard. 1 after Fischer, 2 and 3
after Moosleitner and Urbanek, 4 after Teodor, 5 after Jacobi, 7 after Garbsch, 8 after
Serning, 6, 9 Müller-Wille after Blindheim.
84 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
archaeological finds encompasses several pairs of pincer tongs. The Býčı́ skála cave as-
sembly (5th century BC), containes, iron pincer tongs (35.5cm long) among the sledge
hammers and anvils etc. (Fig. 35: 2; see e.g. Pleiner 1962, 62, fig. 10: 1; idem 1980, 192.
fig. 11.6: 2). Grave 466 at Hallstatt contained another pair of pincer tongs, 22 cm long
(Jockenövel 2001, 16, fig. 3). A contemporary Scythian example comes from Yablonovka,
Ukraine, showing flattened tips; the implement is 16cm long (Fig. 35: 1; see Shramko
et. al. 1969; Pleiner 1980a, 392, fig. 11.6: 1). La Tène period pincer tongs are reported
from an open settlement at Berching-Pollanten in Bavaria (Jockenhövel 2001, 96. fig.
3: 9 and 10). Later finds are more or less sporadic: there were pincer tongs of 25.4cm
length in the Roman hoard of Selz (Ohlhaver 1939, 121, fig. 50, here Fig. 35: 3) as well
as a possible fragment in the smithing assemblage from Sadovec, Bulgaria (6th century
AD, see Welkow 1935; Ohlhaver 1939, 126, fig. 52: 29). One of the three Viking period
objects, that from Alvesta, is made of bronze (21 cm long, see Ohlhaver o.c., 69. Not all
pincer tongs were smithing implements: their function might have been the handling with
glowing charcoal in any fireplace (charcoal pincers mentions Theophilus Presbyter in the
11th century (carponaria) in Schedula diversarum artium VII). Many modern households
have pincer tongs at their chimney fire-places).

Hinged tongs

The principle of this implement is firmly rooted in the two halves joint by a riveted
pivot separating the working part - the jaws - and holding part - the arms (longer part).
The lever effect increased considerably and enabled a favourable grasping power which
could be maintained by various forms of arm-locks (hooks, rings, perforated bands) alle-
viating the smith’s hand. Its invention remains in darkness but the history of effective
hinged tongs is long enough and is connected with the development of ironworking. This
happened during the first centuries of the 1st millennium BC. Hinged tongs were depicted
several times on Greek black- and red-figured vases of the 6th - 5th centuries BC. Beside
the persistently quoted vases (London lekythos and Boston amphora) red-figured repre-
sentations have to be mentioned (Fig. 29), e. g. the Hephaestus riding a monkey and
holding a hinged tongs, a hammer and what could be held for a cool bloom in his hand
(Jockenhövel 2001, 96, fig. 1: 2, after Brommer). Another bowl from Antikmuseum at
Berlin shows a smithing smith working with hinged tongs (ibidem, fig. 2: 1, after Zim-
mer). Iconographical sources do not provide details concerning the adaptation of jaws.
This may be recognized an studied on archaeological finds which can be divided into
several functional types.
In general, universal tongs may be called tools the jaws of which bear pointed tips
enabling a firm grip and preventing the sliding off of the worked piece. At the oppidum of
Manching was found an iron block or bar (No 1498) with traces of such a grasping (Jacobi
1974, pl. 76). Tongs with S-shaped jaws and parallel converged tips occur frequently. In
total they measure 50cm to 60cm on average; the proportion between the length of jaws
and arms varies greatly from 1 : 5 to 1 : 2.5 (the shorter jaws, the more intensive strength
can be developed). Examples are shown in Fig. 37; they cover all periods from the La
Tène up to the Middle Ages and practically all of the European regions. They are depicted
on the reverse side of an Etruscan coin from Populonia or shaped symbolically as brooches
(Fig. 38).
Tongs with specifically adapted or shaped jaw tips include several kinds. Hooked tips
THE SMITH’S TOOLS 85

Figure 35: Pincer tongs. 1 Yablonovka, Ukraine, Scythian; 2 Býčı́ skála cave, Moravia,
late Hallstatt period; 3 Selz, E France, Roman, from a hoard. 1 after Shramko, 2 after
Pleiner, 3 after Ohlhaver.
86 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 36: Hephaestus with hammer and pincer tongs. A bronze stauette in a Berlin
museum (not specified). After Saska and Groh.
THE SMITH’S TOOLS 87

Figure 37: Tongs with pointed and S-shaped universal jaws. La Tène period: 1 St
Georgen, Austria, length 52cm; 2 Kappel, SW. Germany, 52cm; 3 Manching, Bavaria,
17.5cm, arms locking; 4 Golling-Nikolausburg, Austria, the longer example, 71.6cm; 5
Sanzeno, N Italy, 48.6cm. Roman period: 6 and 7 Cáceres el Viejo, Spain, 51cm and 40cm
respectively. Middle Ages: 8 Nordre Besseberg, Norway, locking bar; 9 Tjele, Jutland,
31cm; 10 Kniazha Gora hillfort, Ukraine, 45cm; 11 Mästermyr, Gotland, Sweden, 47cm.
1 and 4 after Moosleitner and Urbanek, 2 after Fischer, 3 after Jacobi, 5 after Nothdurfer,
6 and 7 after Ulbert, 8 and 9 after Ohlhaver, 10 after Kolchin, 11 after Müller-Wille.
88 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
(Fig. 39: 5) suited to clasp smaller flat workpieces as well as those with spatula-shaped
tips (Fig. 11: 3). Sliding sides of worked iron could be avoided by claw or pow tongs
(Fig. 39: 1, 2 and 4). These appeaed as early as the Late Iron Age (Sanzeno) and very
frequently in Roman provinces. Extraordinary long and parallel jaw tips (19.6cm) posses
the unusally long tongs from a rich grave 1 at Vendel in Sweden posses extraordinary long
and parallel jaw tips (Ohlhaver 1939, 135, fig 54: left). The purpose for which operation
they were was designed is not clear.

Figure 38: Smithing tools as symbols. 1 Populonia, Italy, a coin with hinged tongs and
hammer; 2 - 4 fibulae in the shape of smith’s tongs: 2 and 3 after Gaitzsch, sites not
given, Roman; 4 Nowa Boćwinka, Poland, Romano-Barbarian. 1 after Minto, 4 after
Malinowski.

The limbs or just one arm of the hinged tongs were sometimes bent (Fig. 37: 3 and 6)
or ended with ball shaped endings and many of them were equipped with locking rings,
hooks, S-clips or perforated locking bands at arm tips (Fig. 37: 3 and 8). For other
examples see Ohlhaver (1939, figs 28 and 29), reproduced also by Coghlan (1956, 128, fig.
39). A nice pair of tongs with hooked arm locking came from a Romano-Barbarian smith’s
grave at Korytnica, Poland (Malinowski 1951/1952, 259, fig. 2: 1). These systems of arm
locking allow a favourable maintenance of grasping power for a longer period. They are
still in use.
Short sharp-tipped pincers or nippers were observed among Roman implements (Fig.
39: 8); blunt-tipped pliers (39: 9) are known from Viking Scandinavia as well. The find
from Smiss (Gotland, Sweden) comes from a bronzeworker’s grave (Ohlhaver 1939, 131,
pl. 17: 2). These tools were suitable e.g. for the work with wire (cutting, drawing).
Until modern times practice shows that blacksmiths, when working with long bars or
THE SMITH’S TOOLS 89

Figure 39: Tongs with adapted jaws. Tongs with U-shaped and widened jaw tips: 1
Sanzeno, N Italy, La Tène period, length 64cm; 2 Heidenberg-Kreimabch, SW Germany,
Roman, 65.5cm; 3 Cucuiş, Romania, Roman, 72cm; 4 Kolyu-Krupnik, Bulgaria, medieval,
53cm. Tongs with hooked jaw tips: 5 Kolyu-Krupnik, medieval, 48cm; 6 Wlostowice-
Pulawy, Poland, Romano-Barbarian, 36cm. Tongs with long jaw tips: 7 Vendel, Sweden,
early medieval, 116.5cm. Nippers: 8 Sremska Mitrovica, Serbia, Roman, 20cm. Wire
pliers: Bygland, Norway, early medieval, 23cm. 1after Nothdurfer, 2 and 9 after Ohlhaver,
3 after Iaroslavchi, 4 and 5 after Chengova, 6 after Kokowski, 7 after Müller-Wille, 8 after
Popović.
90 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
rods, do not use tongs but hold the cold ends with their hand because the heat gradient
from the hot tip allow it (see also the remark by Sim, 1992).

Cutting and piercing tools

The manufacture of hammers and tongs require implements necessary to make shaft
and rivet holes. Chisels, which were rectangular in section, or drifts and punches with
circular cross-sections served to perforate shaft holes in hammer heads. Implements have
been identified on European archaeological sites which might have fulfilled these tasks,
especially when appearing in assemblages involving smith’s tools (chisels as such were
used in the work of other crafts, too, which worked stone and timber). The circumstances
in which these implements were uncovered do not help very much in thir identification
because other implements were present both in graves and hoards (woodworking tools,
agricultural implements and the like).
Forging practice says that hot-setts with a sharper edge should be applied for cutting
red-hot metal whilst cold chisels’ edges require a more massive cutting part. Smith’s
hammer cutting-edges should be steeled (carburized or with welded-on steel parts). No
metallograpic analyses have been undertaken in recent years but Beck (1884, 544, 666)
reports that two chisels, one from Mainz and the other from the Bibracte oppidum were
equipped with welded-on steel.
Tools with a sufficiently long shank could be operated with the left hand; however,
they might have been shaped as shaft-hole heads (also as cross-edge pieces). Implements
interpreted as smith’s chisels date from the La Tène period onwards (Fig. 40).
Similar operational properties are expected from drifts, punches, and mandrels (Fig.
41) destined to pierce hot and cold metal either in the process of the manufacture of
hammers and tongs or of other artefacts which have to be perforated with holes. Shorter
drifts (10cm to about 15cm) were held in the hand or with the aid of a twisted rod
(setts). A larger Roman circular-sectioned implement can be mentioned from Karabash,
Serbia, 20cm long. It was shafted through a circular hole (Fig. 41: 7). No metallographic
examinations of the working parts and points are available. One would presuppose some
kind of steeling of these stressed parts. Smiths quench their tools repeatedly during work.
It was, without doubt, in the past as well.
Having at their disposal hammers, tongs and piercing tools the smiths were prepared
to use them in both hot and, less often, cold work on a hard cushion, the anvil. This could
have been a stone or adapted stone block which was sometimes still used many centuries
after the beginning of the Iron Age.


Late Bronze Age bronze smiths developed cast bronze anvils, often with roof-shaped faces
and side horns usable in bending operations. The height of these relatively small objects
was some 7cm - 10cm, the protruding horns were 4cm - 8cm long. The more or less
pointed or narrowed bodies were apparently fixed in wooden stems. Some pieces were
ornamented. The geographical distribution according to Ohlhaver (1939, 105 - 111, pl. 2:
4 to 6) and Mohen (1978) covers Switzerland, France, and in the proto-Germanic milieu
in Germany, Poland, and central Bohemia (Velim, Lusatian culture). The ironsmiths had,
then, predecessors of these implements.
THE SMITH’S TOOLS 91

Figure 40: Smith’s chisels. Hot-work chisels with shaft holes: 1 Saalburg, W Germany,
Roman, 10cm; 2 Carichin Grad, Serbia, Roman, 4. 5cm; 3 Bobrichtche fort, Russia, early
medieval, 13cm; 4 Podbolot’e cemetery, Russia, early medieval, 15cm (1 and 3 cross-cut
chisels. Cold-work chisels: 5 Manching Oppidum, Bavaria, 29cm; 6 Selz, Alsace in E
France, 11.5cm; 7 and 9 Bygland, Norway, early medieval, 7.5cm; 8 Bytom Odrzański,
Poland, Romano-Barbarian. Early medieval anvil chisels: 10 Risegjerdet, Norway, 6cm;
11 Vik, Norway, 6cm; 12 Rayki fort, Russia, 12cm. 1, 6 and 9 - 11 after Ohlhaver, 2 after
Popović, 3, 4 and 12 after Kolchin, 5 after Jacobi.
92 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 41: Punches and drifts. 1 Sanzeno, N. Italy, La Tène period, 10cm; 2 - 3 Mag-
dalensberg, Austria, early Roman, 9.8cm and 8.3cm respectively; 4 Sremska Mitrovica,
Serbia, Roman, 17cm; 5 Moosberg, S Germany, Roman, 14cm; 6 Hüchelhoven, Germany,
Roman, 11.4cm; 7 Karabash, Serbia, Roman, 20cm; 8 Gross Raden, Germany, medieval,
13.5cm; 9 Dragoslaveni, Romania, medieval, 11cm; 10 Novgorod, Russia, medieval, 14cm.
1 after Nothdurfer, 2 and 3 after Dolenz, 4 and 7 after Popović, 5 after Garbsch, 6 after
Weinand, 8 after Schuldt, 9 after Comşa, 10 after Kolchin.
THE SMITH’S TOOLS 93

Iron anvils

Anvils made of iron, having been introduced during the Iron Age, proved to be effec-
tive tools, since they offered sharp and hard edges welcomed in shouldering and bending
processes and could be equipped with various depressions, grooves, dies and holes for
making nails and rivet shanks and forged wires. The iron anvils can be divided into two
principal categories: simple squarish iron blocks (block anvils) and horned implements
showing protrusions - the horns on one or both sides of the working face.

Block anvils

They were common through the ages up to Medieval period. The implements were not
large, some 5cm to 14cm inheight and some centimetres square of working face (Fig. 42).
Weights are normally not given in descriptions but two examples from Roman Magdalens-
berg weighed 0.6kg - 0.9kg (Dolenz 1998, 164, pl. 51: W 15 - 16). The bodies are slightly
conical to be set into timbers. With the narrower bottom the shape suggests a pyramid or
frustrum of pyramid with slightly curved edges. One of the widened working face corners
could be perforated with a circular hole for fixing auxiliar tools (swages, cutters) or for
inserting iron rods to forge nails (shaping the shang and upsetting heads). This was es-
pecially the case with Roman block anvils (Fig. 42: 5 to 7) from Einzing and Kreimbach,
the latter having a narrower shouldered circular bottom. The Gallo-Roman tombstone
in the museum of Sens shows Cibeliarius, a master smith, standing at his anvil with nail
hole (Fig. 44, see Esperandieu IV 1911, 12 - 13, No 2761). The Roman or Gallo-Roman
anvils were higher and heavier (height about 19cm - 21cm) and that from Jouars Pon-
chartrain was welded-together from at least three blooms; the anvil face was welded-on
steel, metallographically attested, guarranteeing strenght and certain resilience receiving
the strokes (Rebière et al. 1995, 506 - 507, figs 5 to 9). The working faces of modern anvils
are regularly welded-on of steel and this might be presupposed for ancient and medieval
tools as well. The piece from the Skogar museum, Iceland, represents a specific form of
block anvil, the lower pointed thorn sticks into a stone base (Capelle 1980, pl. XVIII: 1,
here Fig. 42: 9).
The survey should be closed with the remark concerning the Roman block anvils which
were not inserted into blocks but secured against moving by sharp bottomed corner tips
(Fig. 43). Some examples, represented on Roman steles are quoted in the final paragraph
of the anvil section.

Horned anvils

These implements were intended for fine work and specifically for bending rods etc. and
can be categorized as those with one and with two horns protruding from the same level
of working face on both sides. One-horned anvils resemble the letter of L. The horn was
short, about 3cm - 5cm (Fig. 45: 2 or 6) or very long (as long as the shank, e.g. 14cm
(Fig. 45: 5), bent upwards or downwards (Fig. 45: 1 and 3), edged or round-sectioned.
The bodies have narrower base to be stuck into wood. Two-horned anvils have horns on
both opposite sides of the working face so that the are T-shaped (Fig. 45: 7 to 11). The
horns were edged on both sides or on one side only whilst the other was round-sectioned
94 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 42: Iron block anvils. Hallstatt and a Tène periods: 1 - 2 Býčı́ skála cave,
Moravia, HD period, height 9.5cm and 11.5cm respectively; 3 Sanzeno, N Italy, 18cm;
4 Golling-Nikolausburg, Austria, from a smith’s hoard, 14cm. Roman period: 5a - 5b
Jouars-Pontchartrain, France, 21cm (5b section showing the composition from several
blooms, steel face hatched); 6 Heidenberg-Kreimbach, Germany, 21cm (nail-making hole);
7 Kastell Eining, S Germany, 19cm (nail-making hole); 8 Moosberg, S Germany, 15cm.
Medieval examples: 9 Iceland, museum at Skogar, block anvil sticking in a stone base,
27cm; 10 Pastyrskoye hillfort, Russia, size not given; 11 Radovanu, Romania, 24cm. 1
and 2 after Pleiner, 3 after Nothdurfer, 4 after Moosleitner and Urbanek, 5 after Rebière
et al., 6 and 7 after Ohlhaver, 8 after Garbsch, 9 after Capelle, 100 see Narysy, 11 after
Coma̧.
THE SMITH’S TOOLS 95

Figure 43: Block anvils with pointed bases as in Roman reliefs. 1 smithing cupids; 2
Gorzegno, part of a smithing scene; 3 Aquileia, a workshop of a smith and/or locksmith.
1 after Blümner, 2 after Zimmer, 3 after various authors.

(Magdalensberg, see Dolenz 1998, pl. 51: W 14). Some two-horned anvils also had nail-
forming holes and, in addition, edged grooves on their shank serving as dies for forming
thin rods (Roman Einzing, see Ohlhaver 1939, 118, fig. 49: left). The medieval example
from Novgorod, N Russia, was in the equivalent place pierced with three openings serving
as wire drawing holes (Fig. 45: 11, see Kolchin 1953, 58, fig. 17: 2).

Field anvils

Small blocks of iron serving to sharpen cold steel blades are called field anvils (in German
Dengelamboss or Steckamboss). The latter German term indicates that the implement
was fixed in a base. In earlier time it was mostly a wooden block and therefore it belongs
to the family of anvils with narroved bottom parts (Fig. 46: 1). For a piece from Mag-
dalensberg (13cm long, 0.447kg) see Dolenz (1998, pl. 51: W 19). A field anvil, 8cm high
(Fig. 46: 5) has been published from the Polish site of Szaniec, Romano-Barbarian
96 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 44: Sens, France, a Roman tombstone of a smith with a hammer and anvil with
nail-hole. After Esperandieu.

period, see Skurczyński (1953, 43, fig. 2). Other prehistoric field anvils were smaller and
lighter (0.2kg - 0.3kg, height about 6cm). In many cases they have trunnions on their sides
not to sink too deep when beaten (Fig. 46: 4 and 8) or their shank was passed through
by wrapped iron bands. Such objects were relatively widespread in Roman provinces
(Ohlhaver 1939, 36, fig. 10: 6 yo 7). Here we present a find from the Roman fortification
at Sponeck, Bavaria (Fig. 46: 2, see Swoboda 1986, pl. 10: 111) which is relatively large
(16.5cm) and has a pointed base. Field anvils were used not only by smiths but above
all by anybody who needed to maintain blades like sickles and scythes. Parts of a metal
sheet could be adapted on their face as well.
At the end of this abridged survey it is necessary to turn back to the normal block
THE SMITH’S TOOLS 97
anvils of Roman culture, here in terms of fixing their bases. In iconographical sources,
especially on stone reliefs, block anvils are not stuck into a bearing plate but on its suface
being secured at the base against sliding by four pointed corner tips (Fig. 43). Anvils of
that kind make their appearance e.g. on the well known relief with a smithing scene from
Aquileia (e.g. Pleiner 1962, pl. V or in Weissgerber and Roden 1985, 4, fig. 2 etc.) on
the tomb stela from Gorzegno (Zimmer 1982, 112) or on that in the Staatliche Museen
zu Berlin (e.g. Wielowiejski 1975, 218, fig. 58) or on a sarcophagus at Saint Agnan in
France (e. g. Weisgerber and Roden o. c., 15, fig. 20), just to point out some examples.
So far as I know, these anvils are not present amongst archaeological finds.

At this point an unusual object (stray find from the beginning of the 19th century) from
the Slavic hillfort Vratislav in N Bohemia (Fig. 47). A block anvil shank (13.5 cm high)
is pierced with a rectangular hole in which an arm of blacksmith’s hinged tongs is inserted
(the total length is 51cm). The narrower top of the anvil is equipped with a groove (swage
for making thin rods or wire) but when forging the implement must have been put upside
down. Whether it could be used as set hammer is not clear (Pleiner 1962, 174, pl. VIII
and IX). It was a transferable complex of tools belonging to a smith who travelled or used
to be invited to fulfil forging tasks.

Swages and dies

Hand ironwork means the production of individual artefacts. The necessity to provide
forged rods, wires, heads with uniform cross-section or shape leads to applicaton of forge
swages and dies (Fig. 48). Classical upper and lower dies or swages are a rare find. Beck
(1884, 539, fig. 121: c to e) depicts some examples without giving sites; Wedel (1959)
presents a typical pair of top and bottom swages from Pompeii. The lower piece was
destined to be placed in an anvil hole. A similar and very small one (3cm of height) was
found in the Viking Age hoard of Tjele, Jutland (Fig. 48: 5, see Ohlhaver 1939, pl. 16:
2). Other models were not inserted into any device (Villand, Fig. 48: 6).
Rivet snaps which were found in La Tène, Roman and medieval sites (Stradonice,
Magdalensberg, Mästermyr (Fig. 48: 2, 3 and 7) have to be considered as kind of die.
Their use was unavoidable in the manufacture of hinged tools, ship planks etc. Thin
forged wires and rods could be sized in groove dies on anvils as was already mentioned
above.
A certain kind of swages represented, in fact, separate nail irons serving, in a hori-
zontal position, for shaping shanks of nails not using the above mentioned nail holes in
anvils. The method was described by Ohlhaver (1939, fig. 39; iconography: Hausbuch
der Mendelschen Zwölfbrüder Stiftung) e.g. ibidem, pl. 45: 1). The tools from various
historical periods see Fig. 49: 3, 4, 11 - 12. The problem arises, in the identification of
archaeological finds of iron implements with holes, of how to distinguish nail irons from
wire drawing devices which were definitely applied in cold work (see below).
98 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 45: Horned iron anvils. Single-horned anvils: 1 Manching oppidum, Bavaria,
height 12cm; 2 Magdalensberg, Austria, early Roman, 24cm; 3 Mezöbánd (Marostorda),
Romania, 6th/7th century AD; 4 Vik, Norway, 24cm; 5 Staraya Ladoga, N Russia, early
medieval, 12cm, face with horn 13cm; 6 Novgorod, N Russia, 12th century AD, 20cm.
Anvils with two horns: 7 Magdalensberg, early Roman, 25cm; 8 Kastell Eining, Bavaria,
Roman, 29cm (perforated for a nail-hole, the shank equipped with a swage-groove); 9
Heidenberg-Kreimach, W Bavaria, Roman, 33cm; 10 Sadovec, Bulgaria, 6th/7th century
AD, 24cm; 11 Novgorod, 12th century, 19cm (the shank with three wire-drawing holes).
1 after Jacobi, 2 and 7 after Dolenz, 3, 8 and 9 after Ohlhaver, 10 after Welkow, 11 after
Kolchin.
THE SMITH’S TOOLS 99

Cold work operations in iron and steel (and non-ferrous metals as well) were performed
with specific tools applied by smiths (Feinschmiede) and locksmiths or those masters who
needed to achieve more precize dimensions or abrasive surfacing of their artefacts.

Figure 46: Field anvils. 1 Magdalensberg, Austria, early Roman, height 12cm; 2 Sponeck,
Germany, late Roman, about 14cm; 3 - 4 Saalburg, W Germany, Roman, 20cm and 10cm
respectively; 5 Szaniec, Poland, Romano-Barbarian, 10cm; 6 Runder Berg, Germany, late
Roman, 9cm; 7 Joa, Norway, 9th century AD, 8.8cm; 8 Tjele, Jutland, early medieval,
7.5cm. 1 after Dolenz, 2 after Swoboda, 3 - 4, 7 - 8 after Ohlhaver, 5 after Skucyński.

Files

Abrasive work on ferrous (wrought iron, steel) as well as on non-ferrous metals and other
materials like antler, bone, horn was was undertaken with files. Numerous bronze files
(one of iron) were found in the graves at Hallstatt (von Sacken 1868, 89, 155, pl. XIX: 12:
this example was about 18cm long, square-sectioned, the tip round-sectioned to smooth
circular openings). Iron or, better, steel files were necessary when planishing and smooth-
ing metal surfaces and sharpening more massive cutting-edges (axes, chisels etc.) and later
100 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 47: Vlastislav hillfort, NW Bohemia. A peculiar set of blacksmith’s tools: set
hammer/anvil with a die groove, inserted tongs. Presumably part of the equipment of
an itinerant, highly experienced smith. Height of the hammer/anvil 15cm, length of the
tongs 52.3cm in total. After Pleiner.

above all in the locksmith’s work in which precision played a decisive role. The effect of
a file is rooted in densely set sharp cuts, transversal, oblique and cross (Fig. 50: 1 - 7).
Coarse files were cut less than six times per one centimeter, fine tools more than twenty
times (Fell 1997). In the paper cited, Fell presents a large set of pre-Roman files from
England which were metallographically investigated, 17 in number; 7 items were quench-
hardened and revealed martensitic and troostitic phases. Metallographic investigation
on the La Tène file from the Steinsburg oppidum in central Germany has been already
undertaken by Hanemann (1921/22) who identified fine and coarse martensitic structures
of the quenched steel. Surprisingly, the La Tène example from the Liptovská Mara centre
in N Slovakia was made of hard hypereutectoid steel but remained unquenched (Pleiner
1982c, 95, pl. 22: 1 - 5). The majority of archaological files is flat and rectangular -
less frequent are square, half-round, and triangle-sectioned, facetted or round files which
principally come from Roman contexts. The site of Magdalenberg delivered the richest
collection of files showing all the above mentioned cross-sections: 53 items incl. 29 flat
files and 7 saw files (Dolenz 1998, 172 - 176, pl. 54 to 56). Saw files are specialized
instruments for sharpening or topping saw blades with spring set teeth. A special notch
near the handle served as a setting-key. Apart of the cited Magdalensberg finds, Mutz
(1968, 156, fig. 1) investigated Roman files and saw files from Augustodunum-Augst,
Switzerland (Fig. 50: 8 – 9). A half-rounded file was welded-together from two kinds of
steel, softer and hard (microphotographs see o.c., 157, fig. 5).
In the llth century Theophilus Presbyter described the method of the manufacture of
THE SMITH’S TOOLS 101

Figure 48: Dies and swages. 1 Pompeii, Italy, Roman (lower and upper swage); 5 Tjele,
Jutland, medieval (small lower swage, 2.6cm). 6 Magdalensberg, Austria, Roman. Ri-
vetters: 2 Stradonice oppidum, Bohemia, 7cm; 3 Magdalensberg, Austria, early Roman,
6.6cm; 7 Mästermyr, Gotland, Sweden, medieval, 10cm. Dies: 4 Magdalensberg, 7.4cm;
5 Tjele, Jutland, early medieval; 6 Villand, Norway, Viking period, 13.2cm. 1 after Vidal,
2, 5 and 7 after Ohlhaver, 3, 4, 6 after Dolenz.

steel files (sometimes with softer inner parts) in his Schedula diversarum artium (3.XVII
- XIX) cut with sharp hammers or secondarily caburized using charcoal with nitric sub-
stances (horn saw dust, wrapping with leather bands before heating) and hardening them
by quenching in the goat urine or that from a ginger-haired boy. These practices, although
alluding to nitridation, indicate, in fact, the catalyzing role of nitrogen in carburization
processes. The German translation of the Theophilus’ texts presents Beck (1884, 982 -
985).
The finds of files may be also consulted in papers dealing with smithing tools (e.g.
Ohlhaver 1939, 70 - 75; Malinowski 1951/52; Coghlan 1956, 116 - 117, 130 - 132). Coarse
cuts can possibly lead to confusion with a similar tool - the woodworking rasp.

Wire drawing irons

It has been mentioned above that hesitation might be expressed about prehistoric and
102 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 49: Nail- and/or wire-drawing irons.l - 2 Staré Hradisko oppidum, Moravia, 10cm
and 6cm respectively; 3 Châtelet, France, Le Tène period, 18cm; 4 Dombóvár, Hungary,
Roman, 36.5cm; 5 Nord-Roldnes, Norway, early Viking period, 18cm; 6 Sanzeno, N Italy,
La Tène period, 19cm; 7 Bygland, Norway, Viking period, 22cm; 8 By, Norway, 11.6cm;
9 Birka, Sweden, Viking period, 8.8cm; 10 Tjele, Jutland, early medieval, 10.5cm; 11
Opole-Ostrówek, Poland, early medieval, 11cm; 12 Novgorod, N Russia, early medieval,
14cm; 13 Sigtuna, Sweden, early medieval; 14 Staraya Ladoga, N Russia, early medieval,
11cm. 1 and 2 Jacobi after Meduna, 3, 5, 7 - 8, 10 and 13 after Ohlhaver, 4 after Gáal, 9
after Arrhenius, 11 after Możdzioch, 14 after Wolters.
THE SMITH’S TOOLS 103

Figure 50: Files. 1 Sanzeno, N. Italy, La Tène period, 28.5cm; 2 Manching oppidum,
Bavaria, 12cm; 3 Silchester, England, Roman, 25cm; 4 Wymyslowo, Poland, late Romano-
Barbarian, 25cm; 5 Bygland, Norway, early medieval, 25cm; 6 Lejre, Denmark, Viking
period (with wooden handle); 7 Novgorod, N Russia, 14th century AD, 19cm. Saw files:
8 Magdalensberg, Austria, early Roman, 13.9cm; 9 Augst, Switzerland, Roman. 1 after
Nothdurfer, 2 after Jacobi, 3 after Gaitzsch, 4 after Malinowski, 5 and 6 after Müller-Wille,
7 after Kolchin, 8 after Dolenz, 9 after Mutz.
104 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 51: Metal sheet shears. 1 Mainz-Kastell, W Germany, Roman, length 52cm; 2
Sadovec, Bulgaria, 6th/7th century D, 46cm; 3 Romfoghellen, Norway, Viking period,
25.8cm; 4 Mästermyr, Gotland, Sweden, early medieval, 31cm; 5 Tjele, Jutland, early
medieval, 24cm. 1, 3 and 5 sfter Ohlhaver, 2 after Welkow, 4 after Müller-Wille.

medieval irons with rows of holes: nail or wire drawing implements? The drawing of
wire of varying thickness supposes the use of suitable tongs catching the end of the wire
drawn and a device perforated with conical holes of different calibration. The thick
wire enters the larger side of the hole, is thinned and leaves the narrower side of the
opening. Normally, the iron implement is fixed in the vertical position (as illustrated in
the 15th century Hausbuch der Mendelschen Zwölfbruderstiftung). However, the fixing of
many archaeological finds appears as not completely clear. The set of holes is very often
positioned in a groove on one side of the drawing iron (Fig. 49: 6 and 7) or the holes
are connected with a scratch (Fig. 49: 8 and 10). It has to be underlined that all of the
earlier drawing irons were used for the fabrication of non-ferrous wires. Iron wires used
to be forged and formed in various groove-dies, e.g. on anvils (see above).
Drawing plates are very rare among the finds (Fig. 49: 13 and 14). An unusual iron
implement kept in the Ermitage of St. Petersburg was found at Staraya Ladoga in N
Russia (Fig. 49: 14, see Wolters 1997, 207, fig. 1). It is 11cm long, 2cm wide and 1mm
THE SMITH’S TOOLS 105
thick and perforated with 78 holes of 0.2mm to 2mm diameter. Metal thread must have
been drawn through them but the mode of fixing prepares difficulties. One idea is that
the metal threads must have been applied in the weawing of brocade. Sedov (1960, fig.
54:7 depicted a similar fragmentary plate (13cm x 4cm, 5mm thick, with 32 preserved
small conical holes) from the fortified centre Borodinskoye in the Smolensk region.

Metal sheet shears and clips

The metalworkers and smiths helped themselves with shearing metal sheets and used
hinged devices ca 25cm to 50cm long. Instead of jaws (like in the case of tongs) they had
blades (usually 1/3 of the total length). In the modern practice they cut metal sheet up
to 1.5mm thick. These tools are known from Roman times onwards (Fig. 51). The length
varies from 24cm up to 70cm. A massive and long example with short blades was found
at Mainz (Roman camp, see Ohlhaver 1939, 69, fig. 34). In the 6th century AD fort
smithy at Sadovec, Bulgaria) two metal sheet shears with one bent arm each were found.
Several items come from Scandinavian Viking Age sites, incl. the Mästermyr hoard on
Gotland.
By clips are meant small clasping vices consisting of two central-pivotted oblong plates,
some centimeterns long, enabling to grasp and hold minute objects during the handling,
instead of fingers or coarse tongs. Such a clips were in use perhaps not only in the metal
work but at Ošaniči, Slovenia, 2nd century BC, an example was among blacksmith’s tools,
hammers, anvils etc (Gebhard 1991).

Forge spoons

The smith attending his hearth needed implements: a forge spoon for adding fuel and
arrange the glowing charcoal heap, and a sprinkling besom to keep the heap surface cooler
and closed against oxidation and to mainain the hot focus inside; fire pokers and hooks
helped to manipulate with the fire and remove sintered slag cakes from the bellows mouth.
There is little chance to correctly recognize pokers and hooks among various shanks,
bars and rods in archaeological assemblages but spoons are a known category. It is
not without interest that the forge spoons (in German Herdschaufel, Feuerschaufel) occur
among the finds from southern cultural provices of Europe (the Celtic and Roman worlds);
they are practically lacking in the Romano-Barbarian and Viking North. Another fact
is striking: the preserved objects are large and long implements (the anvils are minute
things compared with them) - 60cm to 95cm; the spoon plate or blade itself comprised
sbout 100cm square. Possibly, the smith working at the anvil could reach and influence
the various processes at work in the hearth when necessary. Fig. 52 presents some of
the spoons. Those from Ochtrup and Golling-Nikolausburg show partly twisted handles
which was an effective method of preventing bending. Moreover, the latter object was
equipped with a ring on the handle tip, probably for hanging the tool when not in use.
Sometimes, a shorter handle could be prolonged with a wooden shaft, e.g. the example
from the oppidum at Stradonice (this piece was metallographically investigated revealing
a lined ferritic and pearlitic texture, see Pleiner 1962, 185 - 186, pl. XXIV). The Roman
item vo Kreimbach is socketed - clearly indicating the use of a wooden handle. The item
from Bygland is rectangular (Fig. 52, Fig. 52b).
106 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 52a: Forge spoons. 1 Sanzeno, N Italy, La Tène period, length 58.5cm, plate 10.5cm
x 6cm; 2 Ochtrup, NW Germany, 97cm, plate 20cm x 4.5cm; 3 Golling-Nikolausburg,
Austria, La Tène period, 68cm, plate 14cm x 12cm; 4 Stradonice oppidum, Bohemia,
24cm, plate 6.5cm x 5.4cm; 5 Heidenberg-Kreimbach, SW Germany, Roman 33cm, plate
13cm x 18cm . 1 after Nothdurfer, 2 after Wilhelmi,3 after Moosleitner and Urbanek, 4
after Pleiner, 5 Ohlhaver after Lindenschmidt.
THE SMITH’S TOOLS 107

Figure 52b: A rectangular forge spoon (19) in the inventory of a Viking period smith’s
grave from Bygland, Norway (18: stone). Müller-Wille after Blindheim.
108 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

The tools of the blacksmith and smith, so sophistically developed, spread quickly during
the earlier millennium BC and survived, in the handworking craft, until the present. This
is an admirable phenomenon in the human technical and cultural history.
Chapter VII

SMITHING WASTE

The work of any smith leaves various waste materials. In archaeological context they
often represent the only traces of his presence but in other cases their evidencing value
enable to see the ironworking of the past in its technological and, when data of the quantity
are available, economic aspects. Compared with the bloomery metallurgical wastes, the
influence of the blacksmith’s work on the environment was not yet considered and studied
although the depositing of smithing debris in certain parts of towns might have produced
problems.
The principal waste categories are: fuel residue (pieces of charcoal and dust, occasion-
ally mineral coal, and ashes), then the immediate waste products of the blacksmith’s work
(i.e. oxidized iron particles after heating: the ferromagnetic lamellar hammer scale trace-
able during excavation); globular hammer scale having been produced by using various
fluxes during its removal from the hot metal surface may be classified as belonging to the
most important category of smithing slags (amorphous, nodular, plano-convex cakes and
their fragments, these kinds of waste must be discussed in detail); broken or/and slagged
pieces of the refractory hearth lining and damaged tuyeres and protecting bellows shields;
abandoned or lost fragments of iron artefacts (damaged tools, left over iron scrap). All of
these kinds of wastes may contributed to understand of a particular blacksmith’s work.

Fuel
Up to the beginning 20th century the principal fuel for developing and keeping the heat in
the blacksmith’s hearth was charcoal, mostly purchased as commercional beech (heating
capacity about 7600 cal/kg). The residues of small fragments of charcoal or charcoal dust
have been repeatedly found on the floors of excavated smithies or crushed around the
hearth or in places where it was stored as raw material for immediate need.
Botanical analyses of charcoal from ancient and medieval smithies are still very rare
up to the present day. Box-tree and oak coal have been reported from a smithy in an
early settlement of the 3rd - 2nd centuries BC at Mas Castellar de Pontós near Ampurias
in NE Spain (Rovira 1998, 71) while oak and pine and a very small quantity of beech
and maple were collected from an early Roman long smithing hearth 10009 at the vicus
Eburomagus (Bram in S France), oak and pine charcoal was collected, and beech and
maple charcoal on a very limited scale (Passelac 1998, 136). The heating of iron in a
workshop of a Roman villa at Ramières, France, was carried out in principle with oak
coal (with not significant admixtures of other plants and trees, see Maufras and Fabre
1998, 215, fig. 8; 216; some of the samples were embedded in the slag calottes). A La
Tène period bloomery ironworks with a smithy investigated at Mšec in central Bohemia
represents an interesting example. The smelting slag-pit type furnaces were apparently
intentionally fed with spruce charcoal (Picea excelsa) while the reheating and smithing
work was performed with pine coal (a single piece of beech coal can be discarded, see
Pleiner and Princ 1984, 166 - 167). As for Różański and Slomska (see Pleiner 2000c, 115
- 116) the coal from coniferous trees reacts better at temperatures 600 - 800 ◦ C, while
the beech coal produces more CO at 800 - 1000 ◦ C. In the future it would be worthwhile
looking at the practice of charcoal selection in ancient bloomeries and smithies.
110 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Mineral coal was rerely applied in smithing operations. In the 4th century BC Theophras-
tus knew that some smiths in the Peloponnese used coal (De lapidibus 2.16). Tylecote
(1962, 234, 244; 1986, 225 - 226) quotes examples of coal used in Roman Wilderspool,
Camerton, and Tiddington and other sites remarking that sulphur contents in the min-
eral coal might not be so harmful for heating massive iron; bloomery smelters avoided
it. All kinds of coal contain some sulphur (bituminous coal about 0.3 - 2.7 - 1.75%; the
heating capacity of brown coal as used by modern smiths is 6500 - 7000 cal/kg). Smiths
at the 15th century AD silver mine at Pampailly, France, definitely used mineral coal
from sources of some 20km away, as the wastes on the floor of the smithy at Vernay
showed (Benoit 1997, 57 - 58). The blacksmiths in Bohemia ceased to use charcoal at the
beginning of the 20th century substituting it by Kladno brown coal; since that time they
have not produced any more PCB calottes - the vitrified slag with borax etc.,even when
originally appeared as cakes, disintegrated in few minutes to amorphous crusts.

Fuel ash

This paragraph deals with the charcoal ash which was left after burning out the charcoal.
It remained in the hearth, so far this was preserved, or in various depressions within the
workshop and trampled down into their floors. The importance of charcoal ash is inherent
in the fact that its chemical composition heavily influences the composition of metallur-
gical slag (Crew 2000a, based on experimental work; a pioneering article). For example,
the CaO content in the ash is 90x - 300x higher than in the wood (depending on tree,
eg. pine or beech, or tree part, eg. stem, branches, bark), P2 O5 400x - 500x, MgO 150x -
530x (Pleiner 2000, 215). The same is valid for smithing slags. Unfortunately, analyses of
charcoal ash are rarely available from archaeological situations connected with smithing
work. Buchwald (2002, 22 - 23) provides table 9 showing the charcoal ash composition
of several samples which correspond to what was aid above. As to mineral coal ash, no
analyses were published neither connected with historical nor modern activities.

Hammer scale

The surface of iron or steel, when heated in the hearth, oxidizes. A skin of iron oxides
develops and at lower termperatures (in the range of about 520◦ - 580 ◦ C) consists of
magnetite Fe3 O4 and haematite Fe2 O3 and results in reddish scales; above 575 ◦ C the
oxidized layer is bluish and comprises iron oxide FeO (as wüstite), then wüstite grown
through with magnetite, then magnetite, and finally haematite and form the pellicle. This
means that a part of forged iron is lost: for one heating 3% - 4% has to be considered as
loss, during longer and higher heating even more. The blacksmith is forced to remove these
oxides. The simplest way is to beat the surface of the red hot iron against the anvil - the
oxides crush away forming lamellar small plates up to 1mm thick (the thicknes increases
with the temparature). This is the lamellar hammer scale which especially accumulates
around the anvil and can be dispersed over the workshop floor. It is ferromagnetic and
helps to identify the smithing operations during the field research. Up to hundreds and
thousands grammes of hammer scale (battitures, Hammerschlag or Zunder, okalina) have
been found in certain smithing places or smithies (e.g. Dunikowski et al. 1998, 140 tab.
SMITHING WASTE 111

Figure 53: Hammer scale from a Roman smithy hearth at Nailly, France. 1 Lamellar
scales, 1cm - 2cm. Micrographs of polished sections: 2 bottom: massive magnetite, top:
vitrified layer with white wüstite dendrites, 128x; 3 bottom: magnetite in a black glassy
matrix; centre and top: vitrified layer with magnetite and wüstite isles, 140x. After
Dunikowski et al.
112 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
II). In the paper quoted nice examples of Roman period hammer scale from Nailly, France,
are presented showing the main magnetite layer with a tiny pellicle of haematite (Fig.
53) or with a fayalite cover with dendrites of wüstite (cf. Orengo et al. 2000, 60). Some
scale might have had fallen down to the slag pool near the air-inlet where it could be
embedded in the solidifying mass as such (e.g. Orengo et al. 2000, fig. 25; Mihok et al.
1999c, 124, fig. 6: Fluzin in Mangin etal. 2000, pl. XIV: 7 and 8) or it could form balls
of wüstitic grains (e.g. Mihok et al. 1999a, 25, figs 13 and 16).
Should the smith have welded pieces of iron he was forced to clean the surfaces more
thoroughly from the hammer scale to avoid its presence in the welding seams if possible.
Because of that he removed the oxidized layer with a flux, fine sand or earth or even
with ash - at temperatures about 1100 - 1200 ◦ C the oxides melted. When the smith
used the simple technique of striking the iron against the anvil the molten droplets flew
away rapidly cooling and forming small globules: sometimes they still contained some
magnetite inside and in this case they are ferromagnetic. In other cases they converted
totally to a silica matrix, fayalite and glass, i.e to slag. Allen (1986) describes and depicts
many examples of globular or droplet-shaped hammer scale explained as evidence of
Roman bloom purification at Awre in Gloucestershire, England. Unglik (1991) deals
with ferromagnetic globules (sometimes hollow) and with plate hammer scale from a 19th
century smithy at Bixby, United States, revealing magnetite and ‘paved’ wüstite in their
crystalline structure.
Accummulations of this globular hammer scale can be found around the anvil. They
might have been developed by heavy strokes during forge-welding operations as to some
experiments by Dungworth and Ilkes (2005). When the process of the spreading of fluxes
was intensive, the droplets fell down to the hearth and appeared dispersed or concentrated
creating a slag pool near the tuyere mouth.

Smithing slags

Still in early 1980’s some metallurgists expressed doubts on the existence of any smithing
slag arguing that solely hammer scale could be the waste residual material of smithies
(Piaskowski 1983, 58) despite the fact that also modern smithy hearth fed with mineral
coal produced slag from molten hammer scale, although being dissipated in shapeles
pieces immediately. Archaeological activity after World War II, especially excavations of
settlements, saw increasing finds of slag; at that time, archaeologists were ready to see in
any iron slag the trace of iron smelting.
However, these slags were also found in find circumstances which ruled out bloomery
work, e.g. in settlement agglomerations, towns, larger villages, in castles and monas-
teries. Opinions began to appear which seriously took into account the possibility of
the smithing slag being one of important members of the family of wastes. Later, the
term ‘post-reduction slag’ was introduced to distinguish this kind from the smelting waste
(Serneels 1993, eg. 49 or 157sq.).

The problem of the slag cakes

One of the typical forms which have been steadily discovered in not-smelting contexts
are slag cakes (calottes, culots, Schlackenkalotten) which over time have been denoted as
SMITHING WASTE 113
smithy hearth bottoms (SHB, Tylecote 1986, 173, fig. 113; 1987, 318) or plano-convex
bottoms (PCB, cf. Crew 1991, 32, fig. 3).
The latter term describes the appearance of the objects: a roundish or oval slag piece,
convex at the bottom side, often covered with adhering sand, and flat or slightly concave
with solidified upper surface, sometimes remains and traces of the refractory hearth lining
on one side. When broken, these slags show a porous texture, sometimes layered indicating
that they were created at least in two phases before the smith felt that he had to throw
them out. The size varies greatly: from about 5cm up to 20cm in diameter (the average
being some 9cm - 10cm) and of 2cm - 4cm thickness. The weight oscillates around tens of
grammes, the larger pieces reaching 1kg or so. This depended on time: the smith decided
at what moments the solifiyng slag pool was interfering with the working of the hearth
and the cake should be removed by a hook and thrown out. One or more hours come into
question. The author witnessed the experimental forging of a pattern-welded sword blade
at Eindhoven (1993). A sand flux was used and in about one hour a small slag calotte (ca
5cm in dia.) built up below the tuyere. It is pitty that this experiment was not described
on pages of the Ijzersterk (1994).
The author takes the liberty to draw attention to his personal experience from the
late 1950’s and early 1960’s: in 1958 he visited with K. Bielenin (the excavator of the
renowned Romano-Barbarian smelting region in Holy Cross Mountains in central Poland)
and had the opportunity to see an abandoned 19th - 20th centuries smithy in the village
Zajȩczkowice - with a buried heap of PCB slag cakes in the garden witnessing the activity
of the late smith. The same thing happened some two years later: in the village Klučov,
central Bohemia, once again, in the smith’s garden a deposit of typical ‘calottes’ which
have been produced before the master ceased to use charcoal for heating his iron.
These experiences stimulated the revision of numerous archaeological finds of slag cakes
in question which resulted in the understanding that this type of waste, practically identi-
cal with modern examples, could be traced in ancient urban centres as well as in medieval
towns (even whole layers - at gates in particular) or in rural settlements from any period
of the Iron Age in any country. In the early days of archaeometallurgy a collection of finds
was submitted to chemical bulk analyses and mineralogical examinations carried out by
J. Bartuška. Unfortunately, the results were not presented as a whole; instead, samples
from two localities were published: slag cakes from the 7th century AD settlement at
Epolding-Mühltal, Bavaria (Bartuška and Pleiner 1968) and from the Viking centre of
Haithabu in N Germany (Pleiner et al. 1971). The investigated samples were interpreted
as smithing slags (see also Tylecote 1987, 318 - 319, fig. 8.6a). The Haithabu calottes were
simultaneously investigated by Thomsen (1971a, 103 - 109). Thomsen mainly attempted
to distinguish these wastes from the bloomery slag; he denoted them as ‘Ausheiz- oder
Schmiedeschlacken’, without any allusion to reheating operation (Ausheizen).
Among the cakes investigated by Thomsen were pieces which enabled him to succeed
in the deciphering their formation: the so-called ‘winkelige Schlacken’ (angular slags) are
slag cakes adhering to tuyere or protecting shield air-inlet and indicating that the highest
temperature conditions (about 1200 ◦ C at least) in the hearth were at precisely this point.
The slag pool was not formed at the lowest level of the hearth (as Tylecote presented in
1962 1986, 1987). The convex lower side solidified below the tuyere region in a bed of
charcoal, ash and crushed hearth lining. Thomsen reconstructed the position of the PCB
in Fig. 5 of his paper (1971). Other slag pools were placed in a lower position without
showing any impression from the tuyere mouth or affluence protrusions. This way of slag
114 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
cake formation was accepted later in contributions by Crew (1991, 32, fig. 3) and Serneels
(1993, 51, fig. 41) and was confirmed during some experimental heatings.
From about 1970 the further evolution of the classification of the PCB’s was not ex-
pected: could the cakes be the result of smithing production of iron artefacts or signs of
the purification of iron blooms in its final stage?
The slag cakes from smithing hearths are an important source not only in terms of
recognizing past ironworking techniques but also in the process in evaluation of economic
aspects connected e.g. with circulation of partially worked metal over longer distances.
Therefore, we could not do worse to take a retrospective look at the evolution of different
views. Moreover, such a survey may be useful since with the increasing in number of
those working in this branch of research, the investigators have lost contact with other
scholars working in the same field. Metallurgists, chemists and mineralogists began to
examine this kind of waste by classical chemical bulk analyses and phase investigation
completed by modern analytical and microanalytical methods allowing them to get in-
formation about trace elements (e.g. SEM scanning electron microscopy, XRF X-ray
fluorescence spectrometry, electron probe microanalysis, X-ray diffraction analysis etc.).
Gradually, groups or even schools of scholars were engaged and started to publish the
results of SHB’s or PCB’s investigations and tried to distinguish the refining, purification
or current smithing waste among them.
In 1970 Bachmann presented results of his investigation concerning virtual smithing
slag from Eski Kahta, a medieval Mamluk site of the 11th - 12th centuries AD in Kurdis-
tan, Turkey, and compared it with similar waste from a modern Kurdish smithy in the
same village and documented it by micrography of structural phases. These results also
appeared in his fundamental work (The identification of slags from archaeological sites,
1982, 33, pl. XXVII) which is an introduction to the analytical study of metallurgical
slags in general. Sperl in 1980 produced a monograph devoted to iron slags and their mor-
phological and analytical evaluation (Über die Typologie urzeitlicher, frühgeschichtlicher
und mittelalterlicher Eisenhüttenschlacken). This includes comments on smithing slags
found in the garden of the monastery at Heiligenkreuz, Austria (o.c., 47 - 51). These
works laid the foundations to further studies dealing with historical smithing wastes.
Smithing slag cakes were not only the subject of research of mineralogical institutions
but, exceptionally, in non-metallurgical laboratories: this is the case with the PCB slags
from a Vendel period centre at Helgö in central Sweden. Nearly 300 pieces and 600
fragments were found within building complex 3 which also showed a layered texture or
air-inlet holes on one side. X-ray analyses and Mössbauer spectroscopy were applied by
Danish analysts (Hallinder, Flyge and Randrup 1986).
During the 1980’s the Institut für Geowissenschaften of the university of Mainz be-
came the place where I. Keesmann started a very intensive research work concerning
the post-reduction slags mostly called by him ‘Reduktions-Schmiedeschlacken’. In due
time a special centre called Arbeitsgruppe Archäometallurgie was established and pro-
fessor Keesmann with his collaborators and pupils contributed to what may be called a
mineralogical school producing articles on the archaeometallurgy of iron slags. In 1983
(Keesmann, Niemayer and Golschani) investigation results of one of the oldest sets of slag
was published, i.e. from the Phoenician factory at Toscanos, Málaga province, S Spain
(7th century BC). The PCB calottes in question the mineralogical phases of which are
described in detail, were considered for the first time in connection with the final working
of the steel blooms. The wüstite-rich and layered slag calottes constituted the bulk of
SMITHING WASTE 115
waste from another Phoenician factory of the 7th century BC - at Morro de Mezquitilla
in the same Spanish province (Keesmann and Hellermann 1989). Minute copper prills
were embedded in the Mezquitilla slag cakes which means a certain contamination from
non-ferrous operations in the hearth; this would speak rather for smithing iron objects
than for refining blooms. In 1986 a completed version of the Toscanos and Mezquitilla
examinations was presented which summarized the results (this article is dated from 1988,
however it appeared, in fact, in 1991).
Chronologically very early slag cakes (Hallstatt period D, 5th century BC) come from
the settlement at Niedererlbach in southern Germany (Niederbayern), investigated and
published by Keesmann (1985). The mineralogical composition was extremely heteroge-
neous because of the interal ‘stratigraphy’ of layers which had been formed during several
phases of calotte creation during changing oxidizing and reducing conditions. Relatively
much of the iron (mostly in a corroded state) was embedded showing some MnO-rich in-
clusions. The slags are interpreted as post-reduction waste remaining either after working
of blooms or recycling iron scrap.
Further studies at the archaeometallurgical centre at Mainz concerned Celtic iron-
working. A late La Tène rural settlement at Regensburg-Harting, Upper Palatinate, Ger-
many, yielded some calcium-rich wüstitic/fayalitic PCB slag cakes and their fragments
(Keesmann and Rieckhoff-Pauli 1990). Layering was also observed here and the influence
of charcoal ash (leucite) appeared as non negligable as well as the addition of sand flux
some grains of which did not melt. The waste was a relic of a smithing operation.
Slag cakes of variable shape (plano-convex, convex with a central depression or, vice
versa, with a little domed upper surface) came from the Celtic oppidum of Manching,
Bavaria, and their fragments were examined (Keesmann and Hilgart 1992). Their bottom
sides were covered with sintered sand, sealed by the reducing conditions of a charcoal layer
with molten silicates causing the secondary reduction of some iron prills. The calotte
texture was layered, with different zones of porosity. In one case piece of a fallen-in iron
fragment (3cm long) was found. Two cases yielded particles of non-ferrous alloys. The
MnO content of the samples submitted to chemical bulk analyses was about 0.4%, in
one case 15.11%. Individual cakes may represent either a waste left after refining blooms
(which also were found within the oppidum), the other smithing of objects.
Beyond the northern periphery of the Celtic world, in NW Germany, a Germanic
settlement on the Leine river at Nörten-Hardenberg (periods about LC2/LD1, 250 to 60
BC) produced some dispersed PCB calottes 7 of which were investated by Keesmann
and Heege (1990). The reaction between the hearth refractory lining and slag was again
observed and the use of a sand flux was mentioned. Possible traces of hammer scale come
into question as well. An interesting point is that arsenic containing droplets of copper
alloys were found in the slag, represented by domeykite (Cu3 As). The authors suppose
that the smith also worked with non-ferrous metals and admit the preparation of soldering
substance.
Some of Keesmann’s pupils have published valuable contributions of their own. Kronz
(1997) elaborated a thesis on the detailed phase composition of the non-ferrous and ferrous
slags; he used samples from the Lahn-Dill metallurgical centre in W Germany. In our
context, a special attention has to be paid to large smithing cakes (20cm - 30cm in dia, 5cm
to 15cm thick) from the late medieval smithy at Eschenburg-Wissenbach, site B 85 (o.c.
190 - 197, figs 1566 to 1571; see Chapter IX). They are kalium-rich (‘Leucitschlacken’) for
which fluxes like hearth lining and charcoal ash were responsible. Traces of carburized
116 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
iron are also present. Tap slags indicate the bloom refining in the smithy which is not
surprising in an environment that is permeated with iron smelting ironworks.
The rich archaeometallurgical material from the late Romano-Barbarian ironmaking
site at Kammberg near Joldelund, S Jutland, which brought, after extended excavations,
information of iron metallurgy of the 5th century AD, was treated analytically in the
work by Ganzelewski (2000), another member of the Mainz school. The subject of his
examinations was hammer scale and smithing slag (about 1 tonne was found at the site);
the investigated cakes were of small (ca 9cm in dia) to larger dimensions (12cm to 14cm
in dia, up to 7th cm thick). On the basis of his bulk and trace and phase analyses he
stated all of the parameters that appear in this kind of waste: layerening, additions of
quartz fluxes in two cases and partial existence of pressed out bloomery slag (o.c., 23 -
24, 44, 62 - 64, tab. 5). The verdict admits both the refining/purification of smelted iron
blooms and smithing of artefacts.
The laboratory of the Bergbau Museum in Bochum provided mineralogical and chemi-
cal analyses of slag cakes from the earliest phase of the Roman castra at Xanten - Colonia
Ulpia Traiana in NW Germany (Hauptmann and Mai 1989). They contained, among the
usual crystalline phases, hammer scale, iron/steel particles (0.3 % - 0.5 % C) and a high
silica content which was caused by abundant sand fluxes. They were ordinary smithing
slags except sample D7/9 with high MnO (4.93 %) which originated apparently in the
bloom reheating processes.
Smithing slags as such in form of calottes (represented by about 30% of slag wastes)
were treated in an important article by McDonnell (1991) chemically and mineralogically
analysed around 30 specimens (weighing mostly 200 - 300g) of smithing slags in the
form of calottes from several British sites (and from Burgundy) including medieval York-
Coppergate. Charcoal ash and hearth lining acted in the final composition. In Fig. 2
of his paper there is a section of a typical hearht bottom as it formed below the tuyere
mouth and above the hearth base. In his view, slag inclusions in the iron do not make a
major contribution in the composition of the slag (their presence in the smithing waste,
especially the Mn content is usually very low in comparison with smelting slag). Crew
(1991) undertook a set of experiments involving smelts, billet and bar smithing. He
produced hearth slag when refining the bloom but failed in terms of creating consolidated
PCB slag but he raised no doubt about the process having previously found many PCB’s
in the Bryn y Castell and Crawcwelt smithies in Wales (o.c., 32, fig. 3).
We should not omit the research devoted to PCB slags and hammer scale found at
Kundl-Lus in Tirol near the German-Austrian border. The samples date from the La
Tène period. Maurer (1993) reported on the phase and composition analyses mentioning
considerable sulphur enrichments and low phosphorus contents in comparison e.g. with
slags and irons from the oppidum of Manching.
The research in the field of smithing wastes in Switzerland is stigmatized by the name
of Vincent Serneels (University of Lausanne and later Fribourg) who in course of time
cooperated with Marianne Senn. Serneels published an essential work devoted to the
archaeometry of Swiss iron ores and slags (1993) where he also characterized the smithing
processes and their wastes. In fact, he introduced the notion of post-reduction slag (o.c.
49, 188 sq.) and explained the formation of the slag calottes both during the final bloom
refining and smithing of iron objects. His deductions are based on detailed chemical and
mineralogical analyses and investigations. He underlined the role of certain elements in
the composition of some kinds of ores (Mn, Cr, V) which could influence the slag formed
SMITHING WASTE 117
during the refining of blooms but to a distinctly lower extent when smithing iron and steel
in making artefacts (o.c. Figs 39 and 41). In the latter case the slag may be contamined
by non-ferrous elements when the smith occasionally also operated with bronze and other
alloys. This practically cannot arise during bloom purification and shaping billets from
that iron. Serneels mineralogically, chemically and microchemically analysed a number
of ‘scories en forme de calotte’ from several Swiss Roman sites, larger sets coming from
Avanches-Aventicum (urban crafts) and Marsens (a rural village) which he interpreted as
usual smithing waste (o.c. 162 - 166; 171 - 180, 231 - 222).
In 1996 phase and composition analyses of PCB’s came out in a treatise by the late
Th. Geiger (1996) which were found during excavation in the Roman Oberwinterthur-
Vitudurum, site Unteres Bühl; this was commented by Serneels (1996, 218 - 221). The
matter in question is smithing slags.
Roman urban smithies at Autun-Augustodunum in France produced slag calottes as
well and were treated by Serneels in the monograph by Chardron and Picault (1999, 211
- 213; chemical analyses by A. Ploquin, 283 - 288).
As an appendix to the passage on medieval urban smithies and their slags (cakes as
well) found in Zug-Tugium which was compiled by Senn-Luder (1998) Vincent Serneels
added their chemical compositions (both main and trace elements).
In order to evaluate iron slag from the Swiss localities, both Roman, from Neftenbach
and Dietikon, Marianne Senn applied L’. Mihok, professor at the Technical University
at Košice (Mihok and Senn 1995). The result was that the waste from Neftenbach was
declared as purification slag which was minerallogically different from that from Dietikon
which was a smithing slag formed by an abundant silica sand flux.
The above mentioned institution at Košice has become a research centre which pro-
duced during the last years sets of smithing slag examinations undertaken by L’, Mihok,
the head of the Institute of Metallurgy, and his collaborators. The samples selected
have not only come from Slovakia but also from the Carpathian Ukraine, Moravia, and
Italy. The Dacian slags from Malaya Kopanya were identified, basing on mineralogical
phase examination, chemical bulk analyses and X-ray difraction structural analyses, as
typical smithing slags as well as those from the Carpathian burial mounds of the Romano-
Barbarbarian period (Petrovo, see Mihok et al. 1999a). The same verdict was reached
in the case of the medieval waste from a smithy in the 14th century Moravian castle of
Lelekovice near Brno (Mihok et al. 1997; 1999c). As regards the ‘cake’ 2/5706 it should
be noted that it was not slag but a conglomerate sintered together but not molten layers
of hammer scale (haematite and magnetite).
Considerable attention was paid to smithy slags found during the La Salvia’s excava-
tions at the medieval town of Cencelle, central Italy. Two examination campagnes were
undertaken: during the first, five samples were investigated (Mihok et al. 1999e), during
the second eight (La Salvia et al. 2001). All of them were interpreted as smithing calottes
except CC16 which was claimed to be a refing waste due to the embedded remains of
bloomery slag (elevated MnO content).
The find of a Roman hearth at Nailly, France, inspired the special examination of
smithing waste like hammer scale (see above), slags and also slag calottes (phase and bulk
chemical analyses, see Dunikowski et al. 1998). What is important is the enriching of
the game by external evidence which was available in old French manuals for blacksmiths
from the late 19th and early 20th centuries (Guettier, Lagardelle) in which the practice
of spreading ash and sand fluxes is described. In light of the work with modern industrial
118 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
iron is worth mentioning that soft sorts of wrought metal produced no or little slag whilst
the ‘culots’ might have been expected when working wrought iron of minor quality steel,
and when performing welding operations (o.c., 150).
During the final years of the past millennium another centre of archaeometallurgical
studies emerged to produce analyses of post-reduction wastes: the Institut Polytechnique
de Sévenans, Belfort, France, under P. Fluzin, the head of CNRS Unit UPR A0423. He
was engaged in the evaluation of relevant finds from the ironworks at Val de Gabbia III in
the Alpine Italy, where a large calotte (3.08kg, diameter 17.5cm) was formed containing
droplets of white cast iron (ledeburite); according to facts revealed and in the light of finds
circumstances (bloomeries dating from late antiquity) the waste was declared as refining
slag (Fluzin 1999, 193; 2000).
At that time Fluzin was working hard at the smithy wastes from a number of Gallo-
Roman sites, rural and road smithies in the southern environment of Alesia, France. He
participated in a work by Mangin (et al. 2000, 185 - 291) in the mineralogical evaluation
of numerous slags and calottes, completed by chemical analyses provided by A. Ploquin,
Nancy (ibidem, 271 - 284, 475 - 477). In the introduction of his passage concerning
archaeometry Fluzin pointed out his view on slags formed during post-reduction processes
on iron and steel: either on wastes left by the bloom purification operations (‘épuration) or
smithing of artefacts (‘’élaboration’): both operations yielded slag calottes. In most cases
it is extremely difficult to distinguish them solely by means of archaeometric methods
(see also Fluzin and Leclère 1998), among other reasons because the waste is heavily
heterogeneous in both the crystalline and glassy phases and chemical composition.
Naturally, the culots from purification contain more bloomery reduction slag (which
may be or may be not accompanied with characteristic elemnts from the ore). The slag
formed below the tuyere mouth after manufacturing objects, substantially less contamined
with the ore components but after some time caused problems for the smith’s operating
the hearth and had to be removed. Its texture is more influenced by adding silica and other
fluxes used in the process of hammer scale removing from the surface of the heated iron.
Despite these difficulties Fluzin tried to distinguish, among about 30 investigated calottes,
those which originated in the purification process from those which were the result of
normal smithing operations. Unfortunatelly, all of the samples examined showed similar
mineralogical and chemical parameters. He is inclined to see bloom purification in 8 cases
(Alesia, Blessey-Salmais 104/13 with traces of cast iron, and H 603 and 602), Sombernon L
1/I.2, Flavigny 31/34, Boux 09/40. About 14 samples indicated the ‘élaboration’ (smithing
of artefacts, a half of them coming from the same locaities as above). Five calottes could
have been the product of both working procedures.
V. F. Buchwald (Inst. f. Metallaere, Coppenhagen) had no doubts about the am-
biguous factors after carefully analysing medieval PCB calottes from Norway, Denmark,
Greenland and Sweden (Buchwald 1994; 2001, 52 - 55). In his view all of the calottes
without any hesitation represent evidence of bloom purification.
Evidence that this statement cannot be so strict may be provided by 16th - 17th cen-
tury PCB slags from a smithy at Trosky castle in Bohemia. This workshop served the
fortress and it is hard to believe in the working of imported blooms for that late medieval
or early post-medieval period. The slag was analysed by J. Hošek during his studies at
the Technical University at Liberec (see Prostřednı́k and Hošek 2001). The investigation
offered the usual picture as to mineralogy and chemistry of the examined pieces.
SMITHING WASTE 119

In spite of the fact that smithing slags reveal, in fact, the same mineralogical phases as
smelting slags, their genesis is totally different: they are not a product of the reduction of
ore but, vice versa, the results of oxidation of iron when heated in the hearth. Of course,
variable reducing and oxidizing condition occured in the liquid pool and when solidifying
to cakes. Their silica content did not originate in the smelted ore but in sand which was
spread on the surface of the red hot metal; to a lesser extent, some silica came from the
hearth lining close to the air-inlet hole. On average, the silica content is a little higher
than in the bloomery slags although the proportions certainly considerably overlap.
Contrary to smelting slags the calottes are mineralogically little worlds for their own:
one of the characteristics is their ministratification or layering caused not only by slowing
down or temporary stopping of the supply of fresh molten mass but also by the spread of
internal conditions in the process of cooling. The textures are extremely heterogeneous.
The principal component is the silica matrix constisting of ferrous olivines and some
glass. The main constituent is fayalite Fe2 SiO4 or 2FeO.SiO2 in variable proportions
and cristalline shapes. In some PCB slags an iron silicate iscorite FeII III
5 Fe 2 SiO1 0 was
identified. Both the refractory lining and charcoal ash might have supplied some lime and
magnesia containing minerals like kirschsteinite CaFeSiO4 or monticellite CaMgSiO4 as
well as ferro-akermanite from the familly of melilithes (Ca2 AlSi2 O7 ). The post-reduction
slags in the PCB’s contain variable amounts of iron oxide FeO or, better, Fe1 xO, the
wüstite. This is a product of oxidation and appears in rounded light crystalls in which
another result of oxidation may be segregated - the magnetite. Or the wüstite crystalls in
the fayalite matrix are dendritic, especially when more rapid cooling took place. Magnetite
Fe3 O4 is a member of the spinels like hercynite FeAl2 O4 which also may be recognized in
the smithing slags when the alumina supply (e.g. from the hearth lining) was sufficient.
It has been noted above that charcoal ash had an influence on the internal habitus of
the smithing slags (e.g. Ca, Mg). Another element was kalium in form of potash K2 O
and leucite, a kalium-alumina silicate KAlSi2 O6 which was stated in a number of calottes
as the last crystallization phase of the slag.
It is difficult to offer a general recipe of how to define refining and purification evidence,
within the post-reduction slags and, on the other hand, traces of regular of artefact forging.
The bloom refining requires a gentle pressing out, mostly with a wooden mallet, the slag
from from the pores or cavities of the pasty iron sponge (stage one). The manipulable
metal has to be further purified and forged into the shape of billet or ingot. The workpiece
oxidizes again but in this phase (stage two) the hammer scale has to be removed and fluxes
are spread on the hot surface. Smithing slag develops and creates in the hearth, just bellow
the tuyere mouth, a pool solidifying to a cake, the PCB calotte. In this case it receives
both kind of slag: the remains of bloomery slag still being pressed out from the metal
being purified, and the molten hammer scale and flux. When the original iron ore was
rich in manganese, this lithophile ingredient transits as MnO completely into the slag and
may traced in the purification slag. Similarly the vanadium when it appears in the trace
analysis. Metallic iron can be embedded: when too rich in carbon (hard steel or even
particles of cast iron) the explanation as refining slag is very plausible.
In the usual smithing slags inclusions in the metal do not make any significant contri-
bution to their formation (McDonnell 1991, 29). As to the facts revealed by the natural
sciences, the presence of of non-ferrous metal ingredients speak clearly on behalf of or-
120 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
dinary smithing work; smiths producing iron artefacts very often used copper alloys for
different purposes (plating, inlays, casting-on, soldering) and their slag might have been
contamined with traces of these metals. In a bloom purification hearth this can be hardly
expected.
Let us return to the find circumstances: slag calottes from purification processes can
be hardly expected e.g. in urban and castle smithies of any period where iron artefacts
for daily and military use were manufactured. Moreover, it should be born in mind that
working of imported blooms required additional and not negligable amounts of charcoal
so that in places where there may have been a shortage of fuel the possibility of refining
blooms should be treated very cautiously.
On the other hand, in sites where ironmaking activity is evident (in bloomery iron-
works), then the presence of PCB calottes, very often of larger dimensions, may clearly
indicate the final stages of purification of the smelted iron. To sum up: in the process
of distinguishing slag cakes formed during bloom purification and/or artefact smithing
the find circumstances have to be respected in all cases since the results of archaeometric
investigations can be ambiguous and not able to offer a correct solution. Fig. 54.

However, the huge amount of analytical work undertaken in many centres by experienced
scholars has provided a better insight post-reduction processes in early metallurgy of
iron; simultaneously, it contributed to the improvement and refinement of investigation
methods. As remarked in the introduction to this Chapter, the interpretation of calottes
or smithing hearth bottoms is important for an understanding of the economic aspects in
all periods of the Iron Age, not least in terms of the circulation of iron, the most useful
technical metal.
What happened with the slag calottes when smiths substituted charcoal by mineral
coal and began to use borax instead of sand? It has to be said that a pool of liquid slag
was formed again but it was rather inconsistent dross which disintegrated almost imme-
diately. (In the course of experimental forging of two Celtic sword blades from about
0.75kg heavy commercional steel bars 791g of slag/dross and 172g of hammer scale were
produced in 57 heatings, see Pleiner 1993, 71 - 76).

Amorphous smithing slag

PCB calottes do not occur in majority of smithing wastes. McDonnell (1991, 24) says
about some sites in Britain that the proportion does not exceed 33%). Some of the ham-
mer scale molten by sand flux to a mostly fayalitic mass, dropped down to the hearth
without joining the main pool, the calotte. These amorphous slags (‘scories informes’ in
French papers) are explicitly commented on in connection with smithing slags discussed
from several sites, e.g. the Roman Autun (Serneels in Chardron and Picault 1998), Ale-
sia, Gallo-Roman Blessey-Salmaise (Fluzin in Mangin et al. 2000) etc. By the way, Crew
(1991, 30) did not succeed in producing calotte during his forging experiments; loose slags
were his waste.
Having been found out of smithies and hearths these amorphous slags cannot be iden-
tified as post-reduction waste without mineralogical and chemical treatments.
SMITHING WASTE 121

Figure 54: Smithing slag cakes (plano-convex hearth bottoms, PCB). 1 York-Coppergate,
England, with an adhering tip from the hearth lining; 2 Biberist-Spitalhof, Switzerland,
from a Roman smithy in a villa rustica; 3 Marsens, Switzerland, Roman; 4 - 5 Haithabu,
N Germany, Viking period (4 with an adhering hearth lining and tuyere mouth, 2 two
solidified layers); 6 Blessey, France, from a Roman smithy; 7 Epolding-Mühltal, Bavaria,
7th century AD; 8 Budeč hillfort, Bohemia, 11th century AD (with two solidified layers);
9 Mutějovice, Bohemia, from a 13th century AD rural smithy. 1 after McDonnell, 2 after
Schuzany, 3 after Serneels, 4 after Thomsen, 5 after Westphalen, 7 - 9 after Pleiner.
122 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Iron scrap

When fragments of iron objects or sheets or parts of bar pieces were not used as recycling
material in the forge, they fell into the category of wastes, either they were neglected or
could not be used before the workshop has been abandoned. Let us present just three
examples which represent certain complexes which are not dispersed over excavated areas.
In the Celtic oppidum of Manching a very intensive smithing activity seems to be
developed during La Tène periods C2 and D1a. In pits 791b1 and b3 and 830a/b numerous
fragments of ironwork were found, among them damaged tools and parts of iron bars and
folded sheet (Sievers 1992, 195 sq.). They were originally destined for recycling but
became scrap.
As will be shown in Chapter IX such a scrap was left in smithies. In houses II and IV
at Menzlin, a harbour site of the 9th - 10th centuries AD at the Baltic coast, some knives
and arrowheads were found; in the surrounding area, e.g. in pit V/15 parts of iron bands,
rods, bars and folded iron sheet came to light and were apparently also supposed to have
been used as working material.
In the two 13th century smithies at Mutějovice in central Bohemia, forgotten, lost or
thrown out iron artefact fragments were unearthed: the head of a plough-share, two dam-
aged knife blades, fragment of a spur and a damaged, although once repaired cylindrical
padlock (Pleiner 1969c, Figs, 11 and 16).
Chapter VIII

SMITHING INSTALLATIONS

When working the iron, blacksmiths had to use a couple of important installations and
devices which enabled them to process their tasks. They needed a hearth or fire place,
the bellows and an anvil fixed in the floor of their working place.

The hearths

Occasional cold hammering of small or thin pieces of iron and steel was possible without
any special installation. A solid block of metal (e.g. an anvil stake) or even stone served
as anvil on which the metal was shaped with the hammer. Hot shaping of the metal
required, however, a special device: a bellows-blown fire place, the hearth, as space for
keeping the fuel (charcoal up to the beginning the 19th century) under necessary heat of
700 ◦ C to 1200 ◦ C in at least partly reducing atmosphere. The hearths appear as depicted
in many iconographical sources of the classical antiquity and Middle Ages. What concerns
the archaeological evidence, a considerable number of smith’s hearths was discovered and
mentioned in the literature. Most of them may be accepted as such but in certain cases
the identification can be connected with problems, especially when the the documentation
and data are incomplete or cursory.
Easily to recognize are burnt-red fire-pits dug out in the floor of the working place,
sometimes lined with refractory inner coating, or flagged by stones or tiles. Their fillings
contain charcoal and charcoal ash and smithing iron slag, at least within close environ-
ment, as well as hammer-scale with siliceous components (see above), either in form
of dispersed fragments or as well known cakes (PCB, plano-convex bottoms, or SHB,
smithing hearth bottoms). Sometimes the position of bellows may be indicated by visible
spots projecting in the circumference of the fire-pit or by other traces (see below). These
hearth depressions in floors, potentionally well distinguishable, indicate that the smith
was squatting, kneeling or sitting on a stool at his work as certain iconographical repre-
sentations end ethnological observations show (e.g. the Greek black-figured lekythos from
the British museum, see Blümner IV 1886, 365 - 366, fig. 53 on Pl. VI; Pleiner 1969a,
fig. 8; Roman scenes: e.g. Aquileia, see Zimmer 1982, 186-187, No 123; Vatican, Galleria
lapidaria, see Zimmer 1982,, 181, No 14; the notorically depicted ash urn, now lost, as
in Blümner IV 1886, 372, fig. 61, or Zimmer 1982, 189 - 190, No 26). Interestingly, the
hearths of these sitting workers were elevated.
Fire-pits on substructures or dais were condition sine qua non when the smith worked
standing. Examples of some ancient depictions: Rome, Donatilla catacombs, see Dolenz
1998, 35, fig. 9; Rome, a sarcophag as in Neuburger 1919, 53, fig. 61. Archaeological
situations offer mere destructions of such an installation as heap of stones, clay and often
debris sporadically touched with fire so that any plausible reconstruction of the original
hearth is difficult or impossible. Should be the fire place or fire-pit on a substructure
made e.g. of turf blocks, the traces of the hearth disappeared without any trace.
As to the published reports, different types of hearth can be distinguished in terms of
their shape, dimensions, equippment and construction. Before discussing these features
it seems to be useful to mention modern smithing hearths. Meanwhile the post-medieval
124 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
devices in villages and suburbs were still equipped with stone- or brick-built substruc-
tures, usually leaned against the workshop’s wall and supplied with air from the side, the
recent hearths are parts of cast iron work tables, blown by a compressor from the bottom
of the fire-pit. Decarburization of modern structural steels played in the heating process
practically no role. Hoods for flame and smoke and gas ducting were and are in common
use. The ancient and medieval chimney-like hearths, often reproduced in inconographical
sources were connected with exhaustion systems automatically. The modern fire-pits may
be circular (some 35cm - 45cm in dia) or slightly elongated when longer pieces or rods of
steel were to be heated. Interestingly enough, very long hearths, excavated archaeologi-
cally, belong to earliest features of that kind.

Elongated hearths

The earliest features being explained as smithy hearths were considerably elongated spaces
or depressions (lenght to width more than 2:1). The much discussed example comes from
Palestine: Gerar (Tell Jemme). Flinders-Petrie (1928) discovered and described four
elongated hearths interpreted originally as sword furnaces, dated differently from the
12th to 8th centuries BC. The features, the dimensions of which are not given on Plate
VI (Flinders-Petrie 1928) had erected clay walls (ibidem, 14, pl. XXV; 4 - 5). Later,
they were held for smelting devices (Wright 1939, 461, who suggested the later dating)
and again for forges, and Waldbaum (1978, 59) called them controversary. No slag is
mentioned having been found in the features but a number of iron objects, in the same
stratigraphical level, came from a place north-west of the furnace group (Flinders-Petrie,
o.c. 14 - 16, Pl. XXVI - XXXI. Waldbaum (o.c., 88, note 10) mentions that one of the
hearths was planned to be re-excavated by G. W. van Beek in 1970’s but no information
occured since that time which could throw more light on the Gerar features.
Surprisingly, quite similar hearth model appeared in the opposite western end of the
Mediterranean coast in the Phoenician cultural province. It would be, perhaps, too keen
to look for insemination from the Phoenician centres in Palestine and Syria but the
traffic existed, and the knowledge of the iron technology was transferred to the Iberian
Penninsula by that way during the 8th century BC, at least.
The earliest known elongated hearth comes Els Vilars-Arbeca, a town in Catalonia,
dating from the 6th to 5th centuries BC. A long depression with flat bottom was lined
with refractory material which created low walls on the sides (like in Gerar). At one
end it took a horseshoe-shaped circumvallation supported by two stone plates situated
perpendiculary to the long axis. The interior of the hearth was about 100cm long, 40cm
wide and its depth reached some 15cm (Rovira 1998, 67, 70, fig. 4). The short report
informs about charcoal and ash in the filling but not about the slag. Nevertheless, there
are exact counterparts from a coastal urban site at Lattes (Hérault, S France) Fig. 55:
1 - 2, dating from the 4th century BC. Hearth FR 887 in an open air smithy (ilot 1)
was of the same rectangular shape, 95cm long, 26cm wide and 15cm deep, with a stone
plate closing the southern tip, lined with lateral low clay walls 8 to 15cm thick (preserved
height 10cm). It was filled with charcoal and slag (Lebeaupin 1998, 89 - 90 figs 7 and
8). In a stone-walled smithy in the same site (ilot 4-sud) a long hearth of the identical
type was uncovered (FR 775), 95cm long, 25cm wide, equipped with clay side walls and
a flat stone blocking at the southern end. Moreover, at the wall adjacent to this feature
an air-inlet was identified where the bellows must have been situated; charcoal, and iron
SMITHING INSTALLATIONS 125
slag were found in the filling (ibidem, 92, Fig 10 and 11). Later, in the early 2nd century
BC, another metalworking shop (iron and bronze) has been excavated at Lattes. This
was room 1 in a building in the ilot 2 which was equipped with 4 hearths the largest of
which in the centre was elongated (F 829, 72cm x 18cm, 16cm deep). This, and the three
other smaller installations were simple depressions in the floor and did not resemble the
characteristic hearths described above.
No details are known about a long burnt-red hearth, depicted in the Rovira’s article
(1998, Fig. 3; 70cm x 2cm) which was found at Emporion-Neapolis (Ampurias), a Greek
colony on the Catalonian coast. A different type of an elongated forge comes from the
urban smithy at Castellar de Pontós in the same region which dates from he 3rd - 2nd cen-
turies BC. This hearth (FR 262) measuring 80cm x 30cm, partly sunken in the floor, was
flagged with flat stones. In the centre a circular depression, fired and vitrified, contained
much of charcoal and globular hammer-scale (Rovira 1998, 67, 69, 72. Fig. 5).
Elongated hearths serving the smiths in towns, villages and estates, among other types,
were used in the subsequent Roman civilization as known from the same geographic area
and from other provinces; they are believed to be useful when heating long iron bars
(which circulated as trade ware in the western and northern Mediterranean; Feugère
and Serneels 1998). An extraordinary exemple may be presented from Bordeaux (1st
century AD, see Leblanc 1997). The excavated area of ca 20m x 75m yielded fife phases
of smithing activities (slags, hammer-scale) characterized by superlong burnt-red hearths
(lenght 80cm to 450cm, width 15cm to 24cm, depth 12cm to 25cm), about 40 in number.
According to Leblanc the wheelwrights were at work at that place and heated long rods
for iron tyres (and used circular hearths for further handling the tyres as well).
In fact, the problem is with the circulation of air and distribution of the heat in long
hearths. The case of Bram (Vicus Eburomagus in Aquitany) offers an interesting situation:
in the smaller of the two long red-burnt hearths (No 1009, 185cm x 25cm, 25cm deep) a
plano-convex smithing slag cake was found in situ near the centre of the eastern lateral
wall (Fig. 55: 3). As it is known that the PCB-cakes were formed just below the tuyere
mouth, it is difficult to understand the blowing of such long features.
Tylecote (1986, 163 - 164, Fig. 106a) discusses an elongated clay-lined smithing hearth
of a figure-of-eight shape (200cm x 70cm from Wilderspool (the 3rd - 4th centuries AD)
with a tuyere hole pointing from a side to deeper end of the feature. He considers the
use of it when smaller iron objects were heated at that spot; larger objects required the
use of the whole interior, blown by supplementary bellows. The Wilderspool feature and
the stone-lined hearth from Tiddington, found early in the 20th century by May and
Fieldhouse have been originally interpreted as smelting furnaces.
Two kinds of hearths both elongated and circular (40 in number), with hammer-scale
and PCB-slags, were announced from a rural site at Baudecet, Belgium (Mathieu et al.
1994), without details.
In other cultural provinces, oblong hearths appeared rarely and detailed descriptions
are lacking. Let us mention the smithy in the Celtic oppidum of Závist near Prague
(2nd century BC, see Drda 2000) or from the Dacian Sarmizegetusa-Gradiştea Muncelu-
lui (Daicoviciu et al. 1953). From the Middle Ages no clearly eelongated devices are
known until now. A strongly eliptical hearth worked in the smithy at Sarkel-Belaya
Vezha, Russia, where the fire-pit dug into the workshop floor measured 100cm x 50cm;
what is important in this case is that at one of the longer sides a twin-tuyere was placed
with remains of bellows (wood and leather, see Sorokin 1957).
126 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Circular and oval hearths

Elongated hearths were apparently functional, destined for heating longer rods or bars
of iron. In all of the cultural provinces just mentioned they were used along with other
types, sometimes in one of the same workshop. Circular or slightly oval devices play the
most important role since, filled with charcoal, they warranted a manipulable regime of
heat and to a great deal a controllable atmosphere: oxidizing near the tuyere mounth,
reducing in the opposite part or below the prodruding tuyere (Tylecote 1962, 245, fig 63;
see the experimental work by Weiland and Bunge 1991, 583, fig.6: 4).
Circular hearths were functional as well: they suited to to heating of shorter iron blocks,
lumps and small piled faggots. What is striking is the variability of their diameter, from
30cm up to 150cm or dimensions between 25cm x 35cm and 70cm x 80cm. Preserved
depths varied from 15cm up to 40cm. Unfortunately, among hearths from about 30
European sites the most part lacks any detailed description or documentation (the profiles
and sections are presented rarely). They might have been dug into the soil of the workshop
floor, and burnt-red, or lined with refractory clay, tiles or stones. At one point of the
circumference an air-inlet from the blowing apparatus was situated; it might be practiced
either over the hearth rim, or through the soil or ring through a block tuyere or protecting
shield. When the shape of the fire-pit appeared as oval, in much of the cases this might
have been effected by longer use, cleaning, relining; it is difficult to recognize a planned
constructive intention.
As usually, this category of items comes from isolated finds from archaeological rescue
operations as well as from relevant complexes: smithies. What follows, is the presentation
of some examples, roughly in chronological order. The rest of circular and other hearths
is shortly described in Chapter IX devoted to smithies.
The earliest circular hearths were found in smithies of the Greek colony of the 8th
century BC on the island of Pithekoussai, now Ischia, at the western coast of Italy (Klein
1972, Fig.@). Unfortunately, the circular features of about 70cm in dia. are not described
and depicted in detail. From European examples dating to earlier historical periods
(Hallstatt, Early La Tène) deserve attention those connected with iron smelting. At a
fortified site at Waschenberg, Austria, the ironworks comprise three smithies labelled as
M, R, and S. These were equipped with circular hearths (about 30cm in dia.) with stone
setting in the outer ring (Pertlwieser 1971, 58 - 62, Fig. 7: 1).
Some circular hearths were discovered in course of a rescue dig at Celle-sur-Loire,
site ‘Bois du Jarrier’ (Orengo et al. 2000, 47 - 48, Fig. 3 to 5). Hearth 1 (diameter
50cm, 15cm deep) and hearth 2 (diameter 75cm, 15 cm deep) were apparently lined with
refractory clay; one fragment, slagged in its lower part is pierced what was an air-inlet
hole. Either it comes from a partially elevated clay rim of the hearth (according to the
authors’ reconstruction) or from a clay shield protecting the bellows. The hearths were
surrounded by debris: lamelar as well as globular hammer-scale and PCB slag cakes. As
to the amount, the operations are interpreted as small-scale reheating activity on the site,
which dates from the Late Hallstatt/Early La Tne period. Smaller and larger circular
smithing hearths were discovered in the pre-Roman Greek colonies, like at Martigues
(Bouche du Rhône, near Marseille) or in workshops within the ramparts of European
Celtic oppida (e.g. Mount Beuvray-Bibracte with eleven small circular hearths within one
workshop, or Rheinau in Switzerland). No detaled data have been published until now. A
SMITHING INSTALLATIONS 127
remark should be devoted to circular burnt-red features at the Carinthian Magdalensberg
(late 1st century BC). Formerly they were classified as smelting or roasting furnaces,
e.g. those from the T/O area. Hearth R 1 and R 2 (100cm and 180cm in diameter)
were combined, on southern sides, with squarish pits with stakes in corners, filled with
charcoal, which were denoted as ‘Bottich’, i.e. a trompe-like installation on the principle
of air compression caused by falling water (Dolenz in: Straube 1995, 149 - 151; idem 1998,
16 - 17, figs 1 and 2). However, the system of water supply is not explained in detail. Close
toa 1st century AD domed bloomery furnace found at Minepit Wood (Sussex, England)
a hearth was excavated in the westward vicinity which consisted of broken slag plates
over a burnt-red flat bottom (diameter 40cm, depth some 8cm); the norther perimeter
showed traces of clay lining. J.H. Money (1974,9, figs. 16 and 17) wrote that H. Cleere
considered that it was used for smithing. No details are added. The photograph (Fig.
17) shows two post-holes at the southern part of the hearth which are not depicted in the
plan. One is a little puzzled over the smithing/smelting hearth discovered in the 1950’s by
G. Jobey at Huckhoe in Northumberland. This was a rock-hewn semi-circular depression
running at the frontal side obliquely to the floor surface (50cm x 60cm, 25cm deep at the
opposite side where a stone desk protected the bellows, according to the reconstruction by
Jobey. The hearth was filled with slag, hammer-scale and coal while coal and charcoal are
described to having been dispoersed arount the frontal working place. From the filling
came out a fragment of piled and welded-together low carbon iron, and from the area
a secondarily carburized tanged knife. Tylecote (1962, 232-233) treated the feature as
a native workplace of the 2nd century AD but in the second edition of his book (1986,
141, Fig. 85) he refers to new radiocarbon data giving the 7th - 6th centuries BC and
transferred the case from the Roman period to the preceding chapter on the Iron Age.
However, it seems that this re-dating can be premature since from the relevant Jobey’s
report follows that the Huckhoe site was ‘partly reutilised during the Roman period’
(Jobey 1968, 294) and the smithy must not have been in any connection with the Early
Iron Age palisade settlement.
Within the Roman road station at Kriftel, W Germany, a smithy hearth was discovered
between two stone-walled houses. Its shape was ‘pear-like’, internal dia. 90cm x 70cm,
the rim clay-lined (10cm), opened to a pebble-pave circle; the opening was blocked by a
stone. Abundant iron slags are reported (Schoppa 1967, 102 - 104, figs 1 and 2).
Undisputable smithing hearths were parts of urban workshops at Autun, ancient Au-
gustodunum in France (Chardron-Picault and Pernot 1999, fig. 168). Let us mention
a flat bottom fire-pit 6320 (50cm dia., 50cm deep in the floor of smithy 3-3/2 with a
small forepit and supposed bellows position at one side, or hearth 1287 in workshop 1-2/3
(clay-lined, diameter 67vm, 10cm deep) where, at one side, two tiles were inserted into
the floor: presumably traces of a bellows protecting shield. Similar situation emerged at
Petöháza in Hungary, in a smithy of Roman villa rustica (60cm dia., 30 - 35cm deep),
with stone supports on eastern side. Slag and iron tools are mentioned (Gömöri 2000, fig.
84). An unusual slightly oval fired feature appeared at a Roman rural estate at Horath,
W Germany. In room 7a, leaned against an auxuliar stone bank at the souther wall of
the room, a hearth or furnace has been uncovered the profile of which shore the base of a
domed structure (preserved height about 70cm, ground plan 100cm x 80cm). Slag cakes
were deposited on the bottom. Cüppers (1967, 129) calls it ‘Rennofen’, a bloomery fur-
nace. It seems to be unusual to look for such a device in the villa-like complex like Horath.
Possibly, a kind of a smithing hearth would be more plausible interpretation. This expla-
128 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
nation might be supported by Roman iconographical sources where various box-shaped
and chimney-like hearths appear (e.g. Berlin, Staatliche Museen, see Wielowiejski 1975,
218, Fig 58 etc.). Some of the them look even as bizarr, like a house or chapel as on the
well-know basrelief from Aquileia (e.g. Zimmer 1982, 186 - 187, No 123).
The last mentioned installations were so-to-say ‘normal’ in terms of dimensions. From
the Roman times are reported, however, large or even giant circular hearths. In a villa of
the BC/AD period at the Catalonian El Vilarenc a smithy is described (containing slag,
ash and iron fragment) with two hearths: one was a trapezoidal stone destruction, the
other a large circular feature of 136cm in dia. with a smaller circle of stones in the interior
(Pérez-Suñé et al. 1998, 223 - 224, fig. 3, here Fig. 55: 5). Similarly, a relatively large
circular hearth belonged to the equippment of a Roman road smithy at Einöd-Noreia,
Austria. The matter is of a stone paved bottom wit a (damaged) outer cicrce of stones
(diameter 200cm, inner circle 100cm). There are no details on lining; a low stone wall
blocked th hearth from the NW. Iron slags were dispersed around (Schmid 1932). A bit
enigmatic seems to be two groups of not respecting eachh other large circular hearths
at Riom, Switzerland (the 1st to 3rd centuries AD). Apart of smaller features of 67cm -
70cm diameter, large hearths occured in rooms M and 2, six and nine respectively. Some
of them reached 12cm in diameter. They were burnt-red and filled with iron slag and
charcoal (Rageth 1982, 203, Fig.2, 204 - 205; Fig. 4, see below in Chapter IX). These
device maight be used for large iron blocks used during the structure activities. A giant
circular hearth called ‘furnace’ was a part of a fabrica at the military camp of Inchtuthill,
Scotland, in time when the castra was being evacuated and demolished (about 400 AD).
The hearth measured 150cm in dia.; northwards a stone-paved ditch or channel lead out
of the pit (40cm wide). Due to some nails found at the bottom it is believed that the nail,
an enormous sealed hoard of which was fount in the southern part of the fabrica, were
made there (Pitts and St. Joseph 1985, 108, Fig. 22).
A case which escapes all comparisons is the welding installation Corbridge, Roman
Corstopitum, in England. A stone-walled shaft, circular in plan (external diameter 300cm,
internal diameter 180cm) was preserved up to some 70cm. An arch was open to the west,
backed by walled wings. The interior was filled with charcoal and in the centre stood,
still in situ an iron beam consisting of about 20 welded-together blooms (155kg in total).
Tylecote (1986, 157, 163 and 165) asumes that red-hot blooms, heated in the glowing
charcoal have been gradually put onto the beam (a future structure element) and welded
on. This shaft hearth served originally as a lime kiln (Tylecote ibidem, Fig. 97).
This short survey should be completed by two medieval features. At Lebedka, Russia,
a smithy of the 8th - 10th centuries was discovered in which, in front of a pit for a standing
smith, an elevated hearth corpus of dumped clay was installed (180cm x 110cm) In its
centre was a fire-pit (diamater 70cm), red-burnt and clay-lined, with some gravel stones
in the bottom. This fire place was integrally connected with a vertical clay wall (60cm
high); it is plausible that it played a role as a shield protecting the bellows, placed in the
fan-shaped rear space. Unfortunately, no information of air-inlet or tuyere are given in
the report (Nikolskaya 1954, 100, fig. 42: 1). The Mellager site at Trondheim, Norway,
yielded an isolated circular smithing hearth at a position labelled as K3 (McLees 1996,128,
Fig. 10). The bottom of the feature (50cm x 70cm, 30cm deep) was paved with stones. In
the eastern side a feature called ‘bellows-pit’ was adjacent. Remains of clay blocks near
the throat of the hearth were held for traces of a clay superstructure but is more plausible
that they represent the residues of a clay protecting shield.
SMITHING INSTALLATIONS 129

Square-shaped and rectangular hearths

Some of the Roman smithies were constructed by using fired tiles the effect of which
was that the ground plan of the hearth was squarish or rectangular. An example is F61 in
room 1 - smithy at Chartres (1st century AD) where the tiles formed the bottom (50cm
x 50cm) of a shellow hearth depression (70cm x 80cm). The southern side was damaged
by a medieval pit. Some 2m southwestwards there was a small pit, apparently the base
ov an anvil surrounded by spots full of hammer-scale (Pigeau 1994, Figs 4 and 5). Other
nice examples offer the two smithies (No I from the 1st - 2nd centuries AD; No II from
the 2nd - 3rd centuries) in a Roman villa complex at Biberist-Spitalhof, Switzerland. The
earlier smithy yielded a hearth consisting of two flat tegulae (60cm x 80cm) with traces of
clay lining on its western side. The later workshop also was equipped hearth of two tiles
(800cm x 100cm) bordered by tile and clay fragments. This frame was interrupted in the
SE corner where an imbrex secured the air-inlet from a trompe channel system (Schuzany
1994, Figs 3, 5, 14 and 16 - 18).
The hearth in the rural smithy of Gissey 07, France, was differently conceieved: it
adehered to the stone wall of the smithy and framed up by vertically set stone plates.
The rectangular space, with charcoal on the bottom, measured 80cm x 60cm. Hammer-
scale was stated in the northern vicinity of the forge (Mangin et al. 2000, 134 - 136, Figs
26 and 61). Quite similar construction of the hearth was observed in another Roman
rural smithy (3rd century AD) at Ancernant in the Morvan, France. There, the forge
was limited by two stone plates set up vertically in a right angle leaving the southern
side open. The dimensions of the installation were 80cm x 80cm. On the bottom was a
charcoal layer and a PCB cake (20cm dia.) and a perforated (tuyere?) stone (o.c., 258 -
262, Figs.73 to 76). Unfortunately, the original system of function of both installations
escapes, because the existence of smaller fire-pits with the stone-flagged frames is not
possible to prove. A square of a hearth, framed with stones (100cm x 100cm) in a centre
of a smithy is depicted by Galliou (1984, Fig. 7) but no deails are presented concerning
the workshop at a Roman villa at Pont-Croix/Kervenennec in Brittany (1st century AD).
Squarish hearths were built in the Middle Ages as well and the next two examples show
in a better light what was unclear in cases of the Roman features from Gissey and Ancer-
nant. In smithy B85 at Eschenburg-Wissenbach, W Germany (the 12th - 14th centuries
AD) two hearths were discoverd in diagonally oposite corners of the room. In centres of
destroyed stone-flagged structures (about 100cm x 130cm and 140 x 130cm respectively)
were, in both cases, traced burnt-red spots (40cm and 60cm in dia.) signalizing the fire-
pits (Willms 1995, 74 - 77, Fig. 14). Analogous, better preserved, was the hearth of
an early 15th century smithy excavated at Sezimovo Ústı́, S Bohemia (Krajı́c 1993, figs
18 and 19). This was a square 140cm x 140cm, bordered by stone walls (20cm thick)
which surrounded the interior of the hearth on three sides: north, west, and south. It
was interrupted on the eastern side where the free access was supported by by brick-built
plinths (Fig. 55: 6). The bottom of the hearth was covered with crushed small stones
mixed with clay, heavily burnt-red. What is important is that in the centre lowest part of
a fire-pit was preserved (20cm x 20cm). It is obvious that the hearth was elevated above
the floor but its original height is not known. In front of the hearth was a manipulation
space with a pit 40cm x 40cm, possibly the base of an anvil. The spot was full of hammer
130 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 55: Installations in smithies. Hearths: 1 - 2 Lattes, S France (1 FR775, 4th century
BC, 2 FR887, 1st century BC); 3 Bram, S France, F10009 with a PCB-cake in situ; 4
Nailly, France, 1st/2nd century AD, destroyed hearth built of tiles; 5 El Vilarenc, Spain,
1st century BC/1st century AD, from a villa rustica smithy; 6 Sezimovo Ústı́, Bohemia,
early 15th century AD. Water tank: 7 Stöng, Iceland, a medieval stone water tank. Anvil
support: 8 Gastein-Bockharttal, Austria, an anvil holding stone, 16th century AD. 1 - 2
after Lebeaupin, 3 after Pasellac, 4 Dunikowski et al., 5 after Krajı́c, 7 after Capelle, 8
after Cech.
SMITHING INSTALLATIONS 131
scale.
A number of squarish or oblong features interpreted as smithing hearth appeared in
medieval towns, castles or at mines. These do not reveal any traces of a fire-pit, being
just destructions of stones. Originally they were ruins of stone and clay substructures
of elevated hearths, the fire places proper having been demolished. Their location was
mostly in corners or at walls of smithy rooms. The smiths worked standing when heating
their iron or steel. These destruction are hardly to describe; prevailing dimensions were
about 200cm x 100cm. Individual cases will be mentioned in subchapter 7 devoted to
smitheries.
There is the problem with ventilation. Some of the hearths were just sheltered and the
smoke, gases and spark prepared little trouble. Different situation was in closed rooms
where some exhausting system was unavoidable. Unfortunately, unambiguous archaeolog-
ical evidence for the use of some chimney hoods is lacking. As to iconographical sources
a 4th century AD Roman basrelief from Saint Agnan, France, shows two standing smiths
working at an anvil; in the background must be a hearth being blown by concertina
bellows operated by a kind of transmission but what is visible is pyramid-shaped chim-
ney hood (Gruat 2001). Stone-walled chimney hearths are often represented in medieval
iconography concerning the blacksmith’s work.

Other installations
The bellows and bellows protection

The heat necessary to make iron or steel plastic for further shaping is between 700 ◦ C and
1200 ◦ C. To reach such temperatures was possible solely by the use of forced draught to
glowing charcoal by means of bellows. Archaeologically this is signalized by finds of tuy-
eres in relevant contexts. The device proper, made of leather and wood did not survived,
perhaps with one exception: the case of the smithy at the Russian Belaya Vezha-Sarkel
(the 8th - 10th centuries AD) with preseved twin tuyere adjacent to the hearth and con-
nected with wooden and leathern remains of bellows (Sorokin 1957); unfortunately, they
do not allow to recognize the fanning system - the bellows type.
There are two main categories of bellows used in smithies of the past: the skin bellows
(Schlauchgebläse) and the concertina bellows (Spitzbalg). In the first case a skin of animal
(goat, sheep) with tied up leg tips was fitted on a blowing tube, sometimes in pair; the rear
part was open, stiffened with wooden sticks. The operator’s hand, equipped with a strap,
opened the rear slot of one bag which sucked the air, while the other hand compressed it
into the tube. A skilled man was able to work in this style more than two hours. The lack
of archaeological evidence substitutes, to certain extent, the iconography. On painted or
masoned scenes from smithies skin as well as concertina bellows appear. Skin bellows
survived in Europe until Roman times (according to ethnological sources they were used
in Asia and Africa up to modern age). One can see this device e.g. on a Greek sherd (5th
century BC, see Ziomecki 1975, 63, fig. 22) and on several Roman depictions (Weisgerber
and Roden 1985).
The Roman civilization saw the introduction of a concertina bellows consisting of folded
leather between two wooden desks with valves (mooving leather strips), which secured
the pressing and empting the air. Different sizes were in use both for bloomery furnaces
and the smithing hearths. Roman and medieval iconography shows the latter application.
132 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
They could be moved either vertically, with horizontal position of the desk, or horizon-
tally with vertical desks. For Roman examples see e.g the depiction from the Donatilla
catacombs in Rome or on a tomb altars from Rome and Vatican (see Neuburger 1919,
53, fig. 61; Zimmer 1982, 181, No 114; Dolenz 1998, 35, fig. 9). As mentioned above
when discussing the Saint-Agnan Gallo-Roman sacrophagus, a complicated transmission
system mooved vertically set concertina bellows (Gruat 2001). As to medieval represen-
tations a vertical concertina is seen in the ‘Von den handwerkern im gemeyn’ by Roderico
Zamorensis (about 1475 AD; see Flachenecker 1997). Horizontally adapted bellows desks
appear more rarely (a bronze lamp with sitting Vulcanus; the medieval blacksmith in the
illustration in the Welislav bible, Prague, 14th century, etc.).
Another priciple of air-induction was the trompe or ‘Bottichgebläse’ acting due to the
air-pressure of bubbles under a leathern cover above a sort of reservoir.

The bellows protection

The bellows made of organic materials must have suffered by heat and sparks from the
glowing hearth. It was already mentioned that they often were protected by clay or stone
shields or by brick-shape block tuyeres or even by special walls.
Archaeological investigations of smithing places and workshops yielded all of these
arrangements of protection. Sometimes a kind of shield was directly drawn up from
the hearth lining and perforated by an air-inlet for a tuyere being inserted, as certain
fragmentarily preserved examples indicate (Celle-sur-Loire, Hallstatt/La Tène period; see
Orengo et al. 2000, 134, fig. 84:a; Ancernant, a Roman period smithy, see Mangin et al.
2000a, 259). Specially kneeded semi-circular clay shields appeared at Heek, 32cm wide and
19cm high, Romano-Barbarian period (Nikulka 2000), or at the early medieval Haithabu
(fragments, reconstructed size: width ca 32cm, height about 12cm, see Westphalen 1989,
76 - 79, fig. 25 and 26), Ribe (18cm x 16.5cm, see Bencard et al. 1979, 121 - 123, fig. 5)
or Habrůvka-Padouchov in Moravia, the 0th - 11th centuries AD (Souchopová 1995, pl.
23: 2). More frequent were stone shields of similar shape often made of talk ornated with
engraved motifs. They are characteristic for the Viking period in Norway and Denmark.
Fyrkat (Roesdahl 1977, 45) or Reykjavik (Nordahl 1988, 111 - 118) yielded stone-made
examples.
Wall-like stone-set protections may be referred to the smithies at Ulaka, Slovenia (Ro-
man, smithy 1, Schmid 1937, Fig. 5) or at the castle of Trosky, Bohemia (medieval, see
Hošek 2001, 28 and 30, Figs. 1 and 2a). Hearths in form of a chimney, either Roman
or medieval, warranted the bellows protection integrally as seen among iconographical
sources.
As also easily removable bellows mouth protection may be denoted brick-shaped clay
block tuyeres which are known from forge areas in Celtic oppida (Bibracte: Bulliot 1899,
76, 82 - 92, 97; Pleiner 1958, 129, fig. 23: 4 and 5; Hrazany in Bohemia: Jansová 1960,
671, fig. 249: 7; Rheinau in Switzerland, Schreyer and Graf 1995; Manching in Bavaria,
Jacobi 1974, pl. 99). A number of clay tuyeres, better as parts of a smithing hearth
was treated by Orengo (et al. 2000); they come from Gallo-Roman sites westwards from
Lyon, France: Feurs-Forum Segusiavorum, Roanne-Radumna, Toulon-sur-Allier. The
blocks show cubic or ashlar shape about 12cm x 17cm of frontal face, the length is usu-
ally not decipherable, the air-inlet being slightly inclined and conical (the mouth about
2cm). The authors carried out experiments with block tuyere equipped hearths and they
SMITHING INSTALLATIONS 133
feel that opinions on the use of them in the walls of bloomery furnaces need a revision.
In this respect, however, they are not aware that block tuyeres were found still sticking
in the arches of smetling furnaces. Block tuyeres were evidently used in iron making as
well as forging. Early medieval block tuyeres (the 6th - 7th centuries AD, the Vendel
period) were found at the Swedish centre at Helgö (Madsen 1981) but without relation
to any hearth. Their funnction remains unclear: bloom reheating, forging, bronze casting.

Anvil as smithy equippment

Ancient and medieval anvils as a blacksmith’s implement were treated in Chapter VI.
However, anvils placed on blocks or plinths and fixed near the hearth some one or two
steps in distance which enabled the rational swing of arms represent an installation of
the smithy. In workshops excvated only traces of anvil position could be observed. Small
holes in the floor could be remains of wooden blocks having been set in, or stone anvils
as such are reported in many cases. A stone fixing an iron anvil has was uncovered at the
silver mine of Gastein-Bockhardttal in Austria (Cech and Wallach 2000, 117, Fig. 6).Iron
anvils were usually fixed on wooden blocks sticking with their sharp corners on shanks
on or in its upper surface. The blocks, when wide enough, could rest on the floor and
in this case no hope exists to find out their place. Or the block could be sunk into the
floor and left relevant depressions. These variants are shown in numerous illustrations
and iconographical representations.
In an ancient Greek workshop the (iron) anvil is driven just in the floor (the London
lekythos, see e.g. Pleiner 1969a, fig. 8), otherwise the anvil stems or substructures
usually reached to human knees, exceptionally to hips (Rheinzabern stele, Roman, see
Esperandieu VIII 1922, 35).
In smithies revealed by excavations some indications are available in terms of the
distance between anvil and the hearth. In this place some examples are quoted (refer-
ences see Chapter IX). Very close to the hearth (some 50cm) were anvils at Autun 1-2/3
and Ramière (both Roman) as well as in Sezimovo Ústı́ (medieval). At Belaya Vezha-
Sarkel the distance was about 80cm, 100cm at Waschenberg S (Hallstatt), Brandes B68
and Gastein-Bockhardttal (both latter medieval and post-medieval, respectively). Longer
swings (150cm - 200cm) had to do the blacksmiths in their workshops at Autun 1-9/3 a,b
(Roman), Kosel (early medieval, Fig. 55: 8) and at the 14th century castle of Lelekovice.
The distance apparently was given by the style of work and, above all, influenced by the
size of heated irons.

Water tanks

Any smith who operated a glowing hearth served himself with water. Al least he springled
with a besom the burning charcoal heap to develop a crust on its surface in order to keep
the interior red-hot. Then, he had to cool his implements from time to time. Working
with hardenable steel he needed water or some other liquid to carry out the quenching.
Therefore, all of the smithies must have been equipped with containers. Barrels, buckets,
troughs or pottery came into question. When the container was fixed and sunk in the
workshop floor it became a part of installation, a tank.
Among smithies unearthed up to now water receptacles of different kind have been
stated or traced in certain cases which should be mentioned here. Sunk Massalia amphoras
134 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
were found in smithies of the 4th - 2nd centuries BC at Martigues and Béziers in southern
France (Rétif 1998, 106, fig. 108; Olive and Ugolini 1998, 77, Fig. 4, respectively). Stone
tanks appeared in this part of the Mediterranean as well (Mas Castellar de Pontós in
Catalonia (a very early case, 7th century BC after Rovira 1998, 69, 78, Fig. 5) and
in two workshops at Ruscino-Rousillon, 1st century AD (Marichal 2000, 142). Another
stone tank, more than a millnnium later and from far overseas in the North, served at
the hearth in a Viking smithy at Stöng, Island (Capelle 1980, fig. 5, here Fig. 55: 7 ).
A stone-lined receptacle at a wall of the smithy in a Roman villa rustica at Point-Croix
- Kervenennec, Brittany, showed that it collected water from the roof; a lead pipe duct
secured this supply (Galliou 1984, 80 - 81, Fig. 7).
Close to the smithy workshop in house 14 at the Romano-Barbarian Joldelund, Jutland,
5th century AD, a cistern or pool was dug out (115cm in dia., 82cm deep, 840 litres of
water) which is being connected with the ironworking activity (Jöns 1997, 131, 144 - 146,
with further references). A unique find has to be mentioned: a wooden water trough in
the corner of the smithy at Sarkel-Belaya Vezha in Russia, the 8th - 10th centuries AD.
A specific kind of supply was the running water directed to the workshop from a spring
by rills or gutters. The floors of smithies at the silver mines of Brandes, France (13th
century), i.e. huts B 71, B 68, and B 25 (Bailly-Mâitre 1995, 335, Fig. 3) showed narrow
channels; a developed installation of that kind was found in the 16th century smithy of
the Samson mine (Saint-Croix-aux-Mines, Alsace, ibidem, Fig. 2).

Fuel stores and waste deposits

In any smithy a fuel reserve or stock had to be at hand. Charcoal could be stored in
baskets or bags or in shallow depressions in the floor or just as a heap near the hearth
leaving a charcoal-mixed black spot. It is questionable to classify these features as instal-
lations but in several excavated smithies traces of fuel manipuation was recorded. The
same has to be held about the waste deposits, e.g. ash pits. Slags were usually thrown out
of the smithy from time to time when they became incommode. Examples are included
in the following chapter dealing with the blacksmiths’ workshops.
Chapter IX

THE SMITHIES

The above discussed parts of blacksmith’s installations were preserved, to a great deal in
complexes excavated during archaeological investigations, both systematical and rescue
operations. However, these complexes were discovered by chance so that the material
evidence is explicitly fragmentary. It has to be born in mind that during the fully-
fledged Iron Age the blacksmith’s workshops became unseparable features in inhabited
agglomerations: in all of the towns, hillforts, larger villages, castles where the traces of
them should be expected.
In spite of this fragmentariness, the number of smithies rised enormously during last
forty years and new announcement are reported fluently. In late 1950’s and early 1960’s
few smithies, about five, were known among archaeological finds, described ordocumented
(Pleiner 1962, 93 - 95, 103 - 105, 172, 178 - 187).In other cases indirect indications,
mostly the presence of blacksmith’s tools, evoked the presupposition of the activity of a
smithy. Nowadays far more than fifty smithies are reported in the literature. However, the
evidencing value differs: from precisely presented features up to short remarks, comments
and announces.
To obtain a clear picture about these workshops, their characteristic and function
following parameters or data should be at hand: the kind of the structure, the shape of
ground plan and dimensions, data on hearths and bellows position, traces of the air-supply
system (tuyeres, bellows protecting means), indications concerning the presumable anvil
position. Then, the presence of hammer scale and/or of smithing slag (incl. PCB cakes),
its amount and/or weight, charcoal stores and ash waste, and iron objects found (products,
scrap, blacksmith’s tools). The smithies discussed do yield merely a part of information
wanted or needed: in better described cases the amount of information reaches 30% - 80
%. Nonentheless, many fragmentary reports comprise valuable data or details elucidating
the blacksmith’c working conditions which should not be neglected.
The following survey is arranged according to relevant socio-economic and cultural mi-
lieu in various parts of Europe which formed the background of the craft activity: urban
centres in Mediterranean antiquity, in the Roman world (urban, rural, road, estates), in
other settlement agglomerations (Celtic, Dacian, early medieval Scandinavian, German,
Slavic), in medieval towns and villages as well as in castles and specific smithies at mines.
This may characterize, in broader sense, the functional and production role of the work-
shops; direct evidence among features found occurs rarely.

Urban smithies of the early antiquity

From ancient mainland Greece and adjacent archipelago no excavated smithies are known
up to the present time although they have found reflexes in the literature (a smithy as a
warm place for friendly meetings and talks: Hesiod, Works and Days 493m) or as paintings
on Greek black-figured vases: London lekythos, Orvieto vessel).
The earliest complexes discovered come from Greek colonized places. On the island of
Pithekoussai (Ischia) at the western coast of Italy, a metalworkers’ quarter was uncovered
in the Acropolis (now Mezavia Mazzole) of the Euboean town where two or three smithies
136 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 56: Pre-Roman smithies in Europe. 1 Mas Castellar de Pontós, Spain, 3rd/2nd
centuries BC; 2 Martigues, S France, 4th/3rd centuries BC; 3 Lattes, S France, isle 4
south, 4th/3rd centuries BC; 4 Bibracte-Mont Beuvray oppidum, E France, La Tène
period; 5 Mšec, Bohemia, 2nd century BC, smithy at a smelting furnace cluster; 6 Závist,
Bohemia, 1st century BC, reconstruction of a smithy in the acropolis of a Celtic oppidum.
1 after Rovira, 2 after Rétif, 3 after Lebeaupin, 4 after Fichtl, 5 after Pleiner, 6 after Drda.

have been uncovered, dating from the late 8th century BC (Buchner 1969; Ridgway 1973:
detaild by Klein 1972, 36, Fig. 5). Structure III, apparently only sheltered, measured,
2.7m x 2.7m and was equipped with a hearth in the centre, surrounded with iron fragments
and slag and what may be held for hammer scale. In structure IV (6m x 4m) a mud-brick
hearth and two stone anvils were installed. Traces of non-ferrous metalworking (bronze
and lead casting came to light as well. An idea was expressed that Pithekoussai was a
trading post where imported blooms from Elba were worked. Anyhow, worked iron was
needed in that ancient town.
THE SMITHIES 137
Another stream of Greek colonists, the Phocaeans, aimed, about 600 BC, to the coast
of the Golfe du Lyon in the north-eastern Mediterranean, especially to the Rhône river
estuary. Massalia, gradually a foremost center, was founded there. Several daughter sites
were established in the environment and in some of them ironnworking smithies came to
light.
One of them is Béziers (Hérault) where at the Place de la Madeleine, in front of house
1 a cumulation of PCB slags (519 in number but divided into four chronological phases
between 500 and 300 BC) was uncovered. According to Olive and Ugolini (1998, 77) a
forge must have existed which supplied the building activities with necessary structural
iron. Later, during the 2nd an 1st centuries BC, some traces of ironworking, unfortunately
not very disctinct, were found in room 2 of a house in section 1: oxidized iron, sunken
Massalian amphora as water receptacle, elongated hearths 1027 and 1042 in the north-
eastern vicinity (o.c. 77 - 78, Fig. 4).
Much more important features appeared at Lattes, ancient Lattara (Hérault) north-
eastwards of Béziers. Within islet 1 of the excavated part of the town, between streets
101 and 102 an open-air smithy FR 877 was revealed (the 4th century BC), with a
characteristic elongated hearth FR 775 (95cm x 60cm). A slag deposit was in the southern
end of the place where also pieces of scrap-iron were found. Of the same date was the
stone-walled workshop (3.3m x 3.4m) in islet 4-south (between street s 107 and 108)
involving a similar long hearth FR 775 (95cm x 25 cm) and two stone-lined circular
hearths FY 778 and FY 770. Slags, incl. PBS’ and hammer scale were found inside room
3 which was used in the subsequent phase 350 - 325 BC, as well as in the courtyard in
the south. From the later times of the existence of the town of Lattara (the 2nd century
BC) an attention deserves islet 2 between streets 102 and 103B with a workshop 4.3m x
3m equipped with 4 hearths. FS 29 was the largest of them (72cm x 18cm, 16cm deep).
Three others (FS 30, FS 32, and FS 24) at the southern and western walls contained some
iron slags and amorphous scrap and, what is important, traces of bronze, corral and bone
working. The function of the place was interpreted as a small-scale domestic workshop
repairing iron objects and producing ornaments and small things. The 2nd century BC
smithy in islet 36 was severely damaged. Iron slags and bronze fragments were found near
a shallow hearth FS 207 (75cm x 45cm, 20cm deep). Within islet 4-north, between streets
106 and 107, was a smithy (4.6m x 3.7m) which worked in the 1st century BC. Room 1
was equipped with a burnt-red hearth FY 89 which presumably served in bronze working
operations. Crucibles, bronze castings, corral fragments and coins attest the non-ferrous
metalwork. However, iron is represented as well: 125 slags and hammer scale and several
dozens of iron keys. The author of the article (Lebeaupin 1998, 89) is inclined to see in
this smithy, again, a rather domestic workshop.
West of Massalia is situated Saint-Pierre-de-Martigues, an oppidum and, later, a town
where at house C 6 a (western) part of a stone-walled smithy was excavated (the 4th -
3rd centuries BC, preserved dimensions 3.2m x 1.6m). An oval hearth (52cm x 36cm)
belonged to the equippment as well as a sunk Massalia amphore (water tank) and an ash
pit (1.2m x 0.48m). The two stone discs in the north-western corner of the room are not
held for anvil stones. The floor was covered with spots full of hammer scale and what is
called file dust but only one PCB slag cake, 225g, moreover a later intrusion, has been
found. Nevertheless, PCB slags as well blacksmith’s tongs (60cm long) were found in
neigbouring houses and rooms at Martigues (Rétif 1998).
Greeks founded another colony, further southwards at the coastal outlets of Pyrennees,
138 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
known as Ampurias-Emporion, once a small island. In the site of Emporion-Nea Polis an
elongated clay-line and burnt-red hearth was uncoverd, analogous with other featers in
the region, but with no slag or relevat feature resembling a smithy (Rovira 1998, 68 - 69,
Fig. 3).
The eastern coast of Spain witnessed other cultural and civilization influences: Phoeni-
cian and since 814 BC Carthaginian navigators, merchants and settlers founded trading
posts and stations in maritime regions in lands of Iberians (of not well defined origin).
There, also Greek, Etruscan and Italic elements encountered their interests. Remains of
ironworking smithies were found when investigating these sites, some of them of a very
early date.
An isolated chracteristic long hearth FS 284 (100cm x 40cm, 10 - 15cm deep) comes
from Els Vilars-Arbeca in Lleida (Rovira 1998, 67, 70. Fig. 4), dating from the 6th - 5th
centuries. More information is available about the oppidum at Castellruf-Santa Maria
de Marterolles near Barcelona. Room 9 of the building complex investigated housed a
sheltered smithy of the 3th - 2nd centuries BC with a damaged lined hearth (60cm x
40cm). An amount of slags and vitrified fired clay fragments and some iron objects found
nearby might belong to the context (Rovira 1998, 67). Of the same date is the workshop
uncovered at Mas Castellar de Pontós (Gérone). The matter is of a preserved corner (3m x
5m) of house 1 was explained as possible in connection of the building activity. However,
this site yielded, in sum, the largest amount of of smithing slag (in four chronological
phases involving two centuries, the 5th and 4th BC, which throws some doubts concerning
the real process of the activity). Unspecified accumulations of slag and hammer-scale were
recorded, in three cases 20 - 40 pieces of smithing slag calottes in the stone-walled room
(section 9); a stone water tank close to the entrace, a charcoal deposit FS 301 (oak, box-
tree) and a stone-flagged hearth FR 262 (80cm x 30cm) represent the installation of the
interior. PCB slags and hammer-scale are reported as well as two small fire-places FR
264 and 367 which were used for bronzeworking purposes. Among Tartessos ware Punic
pottery was found documenting contacts between Greek and antagonistic Carthaginian
worlds (Rovira o.c. 67 - 69, Figs 2 and 5).
Except of Pithekoussai no specific artisanal quarters have been revealed in towns of
the pre-Roman period in the discussed part of the Meditarranean. The smithies and
metalworkers’ workshops were dispersed in the building cover. Although they seem to be
relatively sufficiently equipped (long hearths supposed to be suitable for heating of long
bars and rods, hinge and pincer tongs (Martigues and Pontós respectively) their produc-
tion cannot be considered as large-scale. Three features at Lattes were explained even as
domestic establishments repairing iron artefacts and producing jewellery. Traces of work
in lead and bronze have been observed in several cases. A temporary smithing activity
at Béziers in front of house 1 was explained as possible in connection with the building
activity. Some doubts may be expressed in terms of the long period (four phases during
two centuries) concerning operations of this kind at the same limited spot. In other sites
unspecified amounts of slag waste were reported, in three cases 20 - 40 smithing slag cakes
are mentioned.

Roman urban smithies


The heir of advanced Mediterranean civilisation became Roman Empire which gave im-
press to a great part of the ancient world in course of more than one millennium. Roman
THE SMITHIES 139
metallurgy of iron achieved incomparable results in European contexts (summarized by
Pleiner 2000c). It has produced quantites of metal for artisanal, military, structural and
domestic use. In Roman towns, villages and villas, and also in hoards were buried (and
uncovered) tens thousand of artefacts - results of blacksmiths’ work, large and small,
technologically sophisticated as well as simple as to their quality. These objects were
manufactured in smithies in town and country. Many of them were traced during archae-
ological investigations of past decades.
Naturally, the picture is, as usually, fragmentary and consists of evidence unevenly
dispersed in space. Apart of iconographically presented smithy scenes and some literary
recorded details a damped light shines e.g. on Italy proper.
Nevertheless, there is the town of Pompeii which is worth of several remarks, despite
the lack of precise information so far recorded in early times of discoveries. In the mid-
19th century Overbeck (1859, 257 - 259) left a remark that at the Herculaneum gate a
workshop of a blacksmith or cartwright (Grobschmied oder Wagner) was stated not far
from the second well in front of a spacious house. Unfortunately only several blacksmith’s
tools like hammers, tongs, and artefacts like wheel tyres and mountings, handspikes are
mentioned; except of a ‘niche for a house genius’ no more details are presented. Gralfs
(1988) speaks of the Herculaneum gate workshop on Via Consularis (R VI 3, 12 - 13) and
adds an information of a furnace and store room with the tools mentioned and hanspikes
over 100cm long which have been tranferred to the Museum (now lost). Another smithy
worked in iron and bronze (toreutics, hinges etc.) extra muros in front of the Vesuvio
gate (uncovered 1893/1900 and buried again). In room 2 on the ground floor consisting
of four rooms 11 chasing hammers should have been found together with lead and iron
bars, smith’s hammers and 55 bronze vessels. In the first storey an undecribed hearth is
mentioned by Gralfs. In a room in the first storey (in front of the latrina) more than 25
iron objets were found. In regio R I 13, 6 another house with evidence of iron and bronze
working was interpreted as small-scale family shop (no more data given). In R I 6, 1 on
Via d’Abondanza a workshop of a faber ferrarius is mentioned (20 m2 ): an elevated stone
hearth, an iron anvil (8cm x 8.5cm, 11cm high) and a hammer as well as locks, keys, hooks
indicate, again, a small family workshop (the 1st storey served as dwelling). In the same
street (R I 6, 12) an ironmonger’s shop with two rooms yielded a hammer, agricultural
implements, 30 keys 4 padlocks and bundles of iron rods. Pilonius Felix aerarius marked
there some bronzes. His smithy remains unknown. The two iron working and the two
combined smithies (one large, the remaining small family plants) at Pompeii produced
iron objects for needs of the town inhabitants.
For further smithy workshops of every class one has to look on Roman provinces.
Serneels (1993, 162 - 164, Fig. 179 and 180) mentions three stone-walled smithies in
Avenches, the ancient Aventicum, Switzerland, namely in islet 19 (1.27kg of PCB slag),
8 E (4.5kg plus bronze waste), and 16 at St. Martin 0.6kg, 33.lkg in total. In addition
amorphous slags were dispersed over various places (17.79kg in total). Mineralogically
analysed samples indicate iron working, only few might be explained as refining slags
form bloom treatment. Little is known about smithies in the Roman urban settlements
and vici like Oberwinterthur (Switzerland) which have yielded anvils, hammers, files and
slag (Doswald 1994b).
Ulbert (1998, 420, Abb. 12) considers a clay-lined hearth at Auerberg, southern
Bavaria, and an adjacent stone block as an open-air fabrica but no details can be evalu-
ated.
140 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
The largest amount of references to urban and other smithies comes recently from
Gallo-Roman France. The built-up cover of Alesia, the successor of the ancient op-
pidum, showed several workshops equipped with hearths and slags (041, 176, H XVIII-001,
F XXVIII 6-014, 408, 409 which are shortly decribed Mangin et al. 2000, 86 - 96, 103,
Fig. 13). An iron block anvil is being kept inthe Museum of Alise-Ste-Reine (ibidem, 105,
Fig. 15).

A capital site in terms of archaeological investigation of smithies is Autun, ancient Au-
gustodunum, where in the area of the Lycée militaire in the SE part of the town has
been excavated. It was entitled artisanal quarter (potters, metalworkers, blacksmiths).
7500m2 were investigated close to the rampart between bastions 1 and 2. The mono-
graph (Chardron-Picault and Pernot eds. 1999) deals with six workshops which should
be shortly characterized in following lines.
Building E in islet (block) A comprised room 3-3 (5m x 6m) with two burnt-red hearths
A and B, ca 40cm in dia., at the northern stone wall. The floor covered a layer mixed
with iron slag and hammer scale. Originally bronze was worked in that room, later iron.
The other smithies were situated eastwards in islet C where the central part in building
C occupied four rooms, and was in phase 2 (about AD 100 or 2nd century AD) connected
with ironworking (preceded by work in bronze in phase 1, AD 40 to 70). The matter
is of room 2-3/2 (subphases a and b, 4m x 3.5m) with a central hearth (80cm x 50cm)
built-up of tiles, and a stone block (anvil ?). Subphase b saw slight adaptations like the
re-installation of the hearth, placing a wooden block which left a small pit in the floor;
in both cases permeated with iron slag fragments (0.5kg in total),hammer scale, ash and
charcoal. The entrance was in the northern corner. The function of linked-up room 2-4 is
not very clear (slightly fired hearth, 40cm in dia.) but it adjoints the two smithies situated
side by side in the south: rooms 2-5 and 2-1. In the first (4m x 6m) the central part of its
floor was covered with a layer of rammed brick concrete in which an oval depression was
visible but no identifiable traces of a hearth could be observed. The amount of dispersed
slag fragments did not exceeded 0.5kg. The floor of the second, neighbouring room 2-1
(4m x 5m), separated just by a wall, showed two floor levels indicating two subphases a
and b. The hearth must have existed in the western corner, however, it left merely some
traces of fire. A limestone block could serve as an anvil. Two amphorae bottoms were
sunk in the floor. The amount of slag reached, again, about 0.5kg. The southern part
served presumably as a charcoal store. Some modification followed in subphase b: another
stone block was installed beside the first, this once with a dent for fixing a metal anvil or
other implement. Hammer scale is reported to be dispersed on the floor. Entrance door
was in the north.
All four rooms seem to belong to one complex and were used, after the occupation by
smiths, for specified kinds of fine work. They are called ‘forges légères’ (o.c. 108 - 110,
Figs. 76 - 79, 208 - 210, Fig. 170).
A little different were two neigbouring smithies in bulding E (eastern part of islet
C). The workshop in room 1-2 (phase 3, the 3rd century AD) had a central circular
hearth (diameter 40cm) with a small bellows protecting wall in the NE side, about 1m
in distance from the base of the anvil. The hearth was covered with debris containing
overheated iron fragments. The eastern corner was full of waste consisting of iron slag
(18kg), hammer scale, lining pieces and charcoal, and six parts of door-hinges. Separated
THE SMITHIES 141

Figure 57: Roman smithies. Urban: 1 - 2 Autun, France (1 building E, workshop 1-


9, phase 3a; 2 building D, workshop 1-2, phase 3); 3 Ulaka, Slovenia, oppidum of the
Iapodi tribe, smithy 1, 1st/2nd century AD. Villae rusticae: 4 El Vilarenc, Spain, about
BC/AD; 5 Pont Croix-Kervenennec, Brittany. 1 - 2 after Chardron-Picault and Pernot,
3 after Schmid, 4 after Galliou.

by a wall the adjacent rrom 1-9 (phase 3, 3rd century AD, 4m x 5m plus a corridor 1m
x 4m, possibly for a staircase) housed a smithy having worked in two subphases, a and
b. Subphase a comprised a central oval hearth 80cm x 50cm with an adjacent pit, about
142 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
2m far from a sandstone block (anvil, 50cm x 50cm with a lot o hammer-scale around).
Between the oval hearth and the anvil there was a rectangular pit. 20cm x 40cm, 20cm
deep, lined with tiles) and bearing traces of heavy heat (another or specific hearth?).
Three shallow depressions in the eastern end of the rooms were filled with wastes: iron
slag (29kg), hammer scale, stones, brick-shaped tuyere fragments. At the NW long wall
a trace of a charcoal deposit/heap covered the floor. Subphase b represents the final
episode of the workshop’s function. Two heating installations adhered to the NW wall: a
circular hearth or furnace, stone- and clay-lined (100cm internal dia), and a rectangular
stone-walled furnace in 3m distance. The anvil stone remained in its place. What is
important: in the centre of the room as well as in a shallow channel running along the
NE wall two agglomerations of bronze coverings for iron (cow) bells were found, and two
complete artefacts of that kind (o.c. 128-131, Figs 106, 109 and 110; 210, Fig. 170). The
bronzeworker reoccupied the shop at the end of its time.
The iron smithies in building E were denoted as ‘forges lourdes’, possibly because
the slag quantity predominated fifteentimes that of the ‘light industry’ in workshops of
building C. The oval hearth in room 1-9/3a was interpreted as suitable for heating longer
bars. Outside, along the rampart some thrown out slags were noted, about 15kg. Among
amorphous pieces the PCB slag cakes appeared as well (o.c., 211 - 213) and mostly globular
hammer scale.
Production of various implements of everyday life and domestic ware was apparently
typical for forges in this quarter, they were situated in ground floors of storeyed houses and
they faced the problem of ventilation, indeed. In other parts of Augustodunum existed
more matllurgical workshops, especially in the southern and northern stripes close to the
fortification; they are not described, see fig. 204, o.c. 267).
No trace was found of late Roman imperial ‘fabricae’ delivering weapons and armoury:
Augustodunensis loricaria, balistaria, clibanaria and Augustodunensis scutaria mentioned
in the Notitia Dignitatum. The same is about other weapon fabricae: at Macôn worked
the Maticonensis sagittaria, at Bourges-Argentomagus Argentomagi fabrica armorum om-
nium. Such armouries were located, by the way, in Italy (Lucca, Cremona), in Lauriacum-
Lorsch, and in the East (Sirmium, Hadrianopolis etc., see Seek 1876, 2nd ed. 1962).


Down south from central Gaul was Gallia Narbonensis, one of the earliest Roman provinces.
At Nemausus (Nı̂mes) west of the Rhône river estuary, a specific find at the ancient Via
Domitia (now Rue Condé) deserves attention: no forge but a rectangular walled tomb (2.5
x 3.5 preserved m, 1.9m deep) was secondarilly filled with smithing slag, hammer-scale,
fired lining pieces (some of them with air-inlet mouths) and reached some 9m3 of volume
(late 2nd century AD). The rescue operation did not allowed to look for the forge proper
(Vidal 1998, Fig. 3). The long hearth from Bram, ancient Vicus Eburomagus, site 133 in
Avenue Lotard, was discussed in Chapter VIII; no workshop ground plans were identified
(Passelac et al. 1998).
At the Mediterranean coast near the promontories of Pyrenees the large hoard of iron
objects of Ruscino (now Château Rousillon) is being described by Marichal (2000). There
are shortly mentioned three smithies, one built-in the former caldarium of a private bath
(slag in the SW corner, a slagged hearth, a water receptacle, rings and keys: late 2nd
century AD). The second, a room of 3m x 7m, was situated near the cardo in a blind
THE SMITHIES 143
alley contained the same kind of waste and a water vessel; finally, the third was in the
rear of house III, room 91; in a depression hiding presumably a destoryed hearth, an
accumulation of slag was found. The same comprised a 10cm thick layer in the western
part of the room (5m x 20m). Among slag there were found iron rings and rod fragments
(Marichal, o.c., 141 - 142).
In an ancient town at modern Carrier d’En Pujol near Barcelona a fragmentary situa-
tion showing elements of a forge (2nd century AD) were uncovered in a building near the
cardo maximus; A thick-walled oblong clay hearth (30cm x 50cm, 20cm deep) contained
80 vitrified ‘nodules’ and 4 pieces of iron slag (Pres Suñé et al. 1998, 227 - 228, Figs 7
and 8).
Bearing north by west of central Gallic regions we find Autricum, modern Chartres,
where in the site of Barbou a part of a smithy has been revealed at the most northern tip
of a block north of the amphitheatre. Two neighbouring rooms 1 and 2 saw the activity of
blacksmiths of the Tiberio-Claudian era (1st century AD). Room 1 was partly excavated:
a flat hearth paved with tegulae (60cm x 60cm, damaged by a medieval feature) was
situated north of what is called anvil bed surrounded by hammer scale and charcoal-
permeated soil. In course of time this room was used as waste deposit the activity having
been transferred to the adjacent room 2. The forge had been removed from room 6 of
the same building block which was finished in the 2nd century. No analyses are available
(Pigeau 1994, Figs 3, 4, and 8). Garnier (1975) reports on a smithy at Chailly en Brie,
ancient Calagum (Seine-et-Marne). There was a heap of slag and ash between two stone
walls in a vicus from the 1st century AD which may represent the ruins of a hearth. A
hammer belongs to this complex. In western France, in the departement Deux-Sèvres, a
part of a Roman town Rauranum (modern Rom) has been excavated recently. Building
1 sheltered two smithies, one in the SW corner (3m x 6m, 2nd century AD) and one in
the eastern end (3m x 3m, 4th century AD). The field situation was badly destroyed but
detailed geophysical measurements identified the position of both workshops on hand of
hammer scale and smithing slag accummulations. The 4th century smithy yielded the
position of the forge hearth (about 1m x 0.5m) and an anvil stone substructure in 1m
distance. Other two smithies could be identified outside building 1.80kg of smithing slag
(44kg in smithy 2 and 11kg in smithy 4 (incl. the PCB’s), as well as fragments of block
tuyeres were found. Oak coal was exclusively used in heating operations (Diedonné-Glad
1997; Dabas et al. 2002.
In the Roman province Britannia apart of the reinterpreted hearths at Tiddington
(Warwickshire, central England) and Wilderspool (Lancashire, northern England) where
no smithy ground plans were published (see Tylecote 1962, 234; 1986, 164, fig. 105).
Mention has to be made of Carlisle (Luguvallium) at the western end of Hadrian’s wall,
and Corbridge (Corstopitum) just in the hinterland of the fortification in the east. In
Carlisle, a Roman fort has been reported in HMS News (47, 2001,2 - 3). An accummu-
lation of hammer scale was found in the via principalis near the place where a fabrica is
supposed to have existed. Corbridge is the site with the huge furnace, formerly lime kiln,
which served as welding device for large iron beams (see Chapter 8). Then, far in the
north, Inchtuthil castra left not only the renowned large hoard of nails but also an area
called fabrica, with a central courtyard (Pitts and St Joseph 1985, Chapter 7, 105 - 111).
The description of acomplex smithy is rather problematic. The large circular hearth was
about 15m away from a rectangular pit filled with slag while the immense sealed nail
deposit was located another 25m to the east. It should be born in mind that Inchtuthil
144 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
base was in the stage of demolition when Roman troops evacuated the region.


Looking eastwards from northern Italy we have to deal with three important localities
where blacksmith’s workshops were discovered. First, in the territory of a Celto-Illyrian
tribe of Iapudes a hill settlement of Ulaka (ancient Terpona, on the Loz/Laas river,
Slovenia) was investigated in the 1930’s by Walter Schmid. A smithy was located in
house 1 (13.4m x 3.4m) with two semi-circular stone-built hearths, one in the centre of
the northern part, another in the SW corner of the room. The hearths were presumably
protected by low stone walls, Fig. 57: 3. Smithy 2 (labelled as A) where a stone-paved
hearth was situated in the centre of a partly excavated room, yielded, according to what
was reported, ‘welding sand’ and several smith’s tools and other iron implements: tongs, a
hammer, shears, an adze, and some cutlery and a number of bronze objects. No iron slag
is mentioned (Schmid 1937, 18 - 21, Figs 3 and 7). Other workshops were not published by
the excavator but Gaspari (1999) presented a concise survey of other Schmid’s discoveries:
further seven smithies with hearths must have worked non-ferrous metals as well (crucibles
and what is called bronze slag). Their activities date from the 2nd century AD (Tiberius,
Marcus Aurelius).
A specific and stratigraphically complicated situation appeared in another urban set-
tlement in Dalmatia: at Blagaj-Maslovare, the site Majdanište upon the confluence of
Japra and Sana rivers, east from Prijedor. The region was an important iron producing
region in antiquity (Pašalić 1954; Sergejewski 1963); Roman slag was quarried and trans-
ported to ironworks at Sisak until 1960. Slags from pre-Roman Maslovare appear in the
subsoil of later built-up covers in the area. The Roman settlement consisted of several
stone-walled buildings from different periods from Nerva (coins) up to the Byzantine era
of Justinianus the Great, i.e. from the 1st to 6th centuries AD (Basler 1977). During the
3rd century a precinct was established with a gate leading to a corridor (about 13m x 5m)
behind which, separated by transversal walls, a room (II) served as a smithy (Fig. 58).
It was in use during the 3rd and 4th centuries (walled by opus incertum, leaving a slag
layer). Production ceased sometime in the 5th century, possibly after raids of Ostrogoths.
Nevertheless, Justinian’s project of the Renovatio Imperii Romani lead to the rebuilding
of the site (basilica etc.), and also the smithy in room II (walled at that time with the opus
spiccatum). Two transversal water channels crossed the room and divided it into three
spaces each equipped with an anvil block (no details given) surrounded by semifinished
iron nails. In addition two other anvil blocks were found, moreover, close behind the room
wall in the western part of the corridor (one of them again with nails). At the northern
wall near the western channel a hoard of ready-made nails (10cm up to 30cm long) was
deposited (Basler, o.c., Pl. II); no number is given. In the final phase of the workshop’s
existence both channel outlets in the wall were sealed. The site was abandoned in the
late 6th century AD, possibly in connection with colonizing Slavs and raiding Avars. The
smithy of Blagaj-Maslovare is, as yet, unique. The report published aims for a detailed
analysis of stratigraphy and mansonry and the installations concerning the ironworking
were mentioned rather in margine, despite of the title of the Basler’s paper.
Ostrogoths were involved in the fate of the fortified system of ancient strongold and
town at Sadovec, near Pleven, Bulgaria, i.e. the site of Golemanovo Kale and that of
Sadovsko Kale. Smithies were found at both: in the first case a room 4m x 5m in the
THE SMITHIES 145

Figure 58: Blagaj-Maslovare, Bosnia. Feature II with Roman smithies producing or


adapting iron nails. After Basler.

fortress casemates of rampart 3 (slag finds, see Uenze 1992); the more important second
smithy was in the ground floor of the pentagonal tower where a workshop with a stone
hearth was identified in which a set iron blacksmith’s tools (horned anvil, hammers, sheet-
shears, tongs, and a bloom) came to light (see Chapter 6; Welkow 1935; Ohlhaver 1939,
146 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
125 - 126, Fig. 52; Uenze 1992, 73 ff.) The date: 6th century AD.

Roman rural and road smithies


Master-smiths did not only exercise their profession in towns but also in the country, in
road stations, trading posts and villages. In fact, the archaeological literature hints at
numerous smithy workshops. However the reports are of different value and sometimes
fundamental elements of evidence appear questionable.
This is the case, for example, with the so-called post station at Einöd in Austria, the
ancient Noreia on the road from Aquleia in Italy northwards to Virunum and Ovilava.
It is marked on the ancient map known as Tabula Peutingeriana. Walter Schmid (1932,
Fig. 95) excavated there a multipurpose stone-walled house, ‘post’, with 6 rooms. The
ground plan was slightly trapezoidal (20.6m - 21m long, 12.75m - 15.9m wide). There was
a circular stone hearth (internal diameter 60cm, see Chapter VIII) in the corner of room 5
screened by a small stone wall behind which the bellows might have been placed. Because
of that Schmid presumed (his words) the existence of a smithy but adds that no finds
elucidating the actual situation were made (the floor was burnt-red). No slag is mentioned
but a flat stone on a mortar bed could have been a bearing plate for a quenching vessel
(o.c., 202). Iron objects found during the excavation of the house included horse-shoe
fragments, chain rings, a knife and a smith’s hammer.
By the Roman road leading from modern Marsens to Riaz (Switzerland, north of Lac
Léman) at the site named En Barras, a vicus has been excavated (1981) of which six
rectangular houses (no one stone-walled) were found along the road. Accumulations of
iron slag bordered the road, the heavy point (80%) being in the northern part of the
excavated area. At one place a gravelled and stamped spot and a large hearth (diameter
150cm) were uncovered on its periphery. A layer consisting of black soil permeated with
charcoal, small pieces of slag and innumerable hammer scale surrounded that heating
installation. All this indicates the existence of a smithy, although it was not possible to
identify the precise ground plan. Most likely a light roofing sheltered the workshop. The
slags, the amount of which reached 2 tonnes, were analysed by Serneels (1993, 170 - 180,
Figs. 192 - 202). Apart of vitrified fragments from a slagged hearth lining, the bulk was
represented by PCB smithing cakes (diameter 5cm to 20cm, weighing mostly between 100
and 700g). Lamellar magnetic hammer scale ocurred in a great quantity. According to
Vincent Serneels the smiths worked imported purified iron, perhaps in shape of a kind of
bar. They supplied the inhabitants and travellers with artefacts they desired during two
centuries and left an immense amount of waste. Occasionally they operated with bronze
alloys as fragments of crucibles and bronze or non-ferrous slags show. At the road net
south of the Lake Neuchâtel, Switzerland, a Gallo-Roman smithy has been discovered
and published recently at Châbles. It finished its production in the 1st century AD. The
matter was of a free-standing complex (6m x 10m) including the room where the smith
was presumably living, a shelter with 2 hearths northwards. In the east were two long
and narrow pits (6m and 5m long, 60cm to 80cm wide) filled with slag and charcoal, of
not clear purpose. The eastern area was covered with wastes, burnt clay, hammer scale,
smithing slag and PCB calottes. A tuyere fragment and iron scrap were found and a
part of a block anvil with nail hole and several chisels. The smithy produced customary
utensils (the authors calculate that about 5 tonnes of iron ware left the workshop) and
recycled iron scrap as well (Anderson et al. 2003, 77 - 174).
THE SMITHIES 147
The station of Kriftel was located on the northern periphery of the Agri Decumates
in Germania (modern Main-Taunus Kreis) and was a burgus securing the road protection
close to the frontier of the Imperium. A rescue dig brought traces of blacksmiths’ activities
(Schoppa 1964, 102 - 103, Fig. 2). Two places are of relevance: one was situated 12m
westwards of a stone-walled tower between two other buildings. Presumably it was just
sheltered. A poorly defined twin installation was in the core: one paved with stones
(called Brennraum), the other in clay (called Heizraum). The feature is discussed in our
Chapter VIII. A lot of iron slag is mentioned. The second installation was to the north,
14m from a two-room stone house ibidem, 103, Fig 1). This was badly damaged but the
author mentions two hearths (Aufbereitungsöfen), an oval, 90cm long, which was replaced
with another one (length 130cm). It was destroyed by hooligans during the excavation
so that the situation could not be analysed any more. Numerous iron nails were found
there, an iron oil lamp and a hook (ibidem, Fig. 6). At Kriftel, iron was worked for needs
of the burgus.
Another type of ironworking in road stations: Gembloux-Baudecet in Belgium (Ger-
mania Inferior) on the road connecting Colonia A.A. (Köln) and Bagaeum (Bavai). This
is a peculiar site. During rescue excavations 42 burnt-red hearths were dispersed in the
surroundings off and within a temple of the 1st to 3rd centuries AD after phases of de-
struction. Twelve were semi-circular (45cm to 60cm dia., 10cm to 15cm of preserved
depth). Thirty hearths were elongated (up to 150m - 200m long, and 50cm deep, with
burnt-red vertical walls). They were dispersed over the excavated area but they also ap-
peared within the temple, chained along its walls creating rows 10m - 11m long. Iron slag
appeared in the occupation layer, several kilogramms of this was filled into a pit south-
westward of the temple. The majority was represented by PCB calottes. It is not possible
to speak about a normal smithy. More likely, specific intensive ironworking operations
took place at that spot producing some long iron objects or structural elements.
In this respect a site has to be introduced here, namely Bordeaux-Cité Judiciaire
(Burdigala in Aquitany). Again, during rescue digs forty super-long and narrow heavily
burnt-red hearths (50cm to more than 450cm long, 15cm to 24cm wide, 12cm to 25cm cm
deep) existed in fife chronological phases during the 1st century AD. Leblanc (1997) sup-
poses that wheelwrights were at work at that place, heating long rods of iron and shaping
tyre strips. Bent tyres were probably treated in specific circular toroidal installations of
140cm dia.
Traces of a destroyed smithy of the 1st - 2nd centuries AD were excavated during a
rescue action in 1994 at Nailly near Sens, ancient Agedincum (Yonne). Overfired tiles and
lime- and sandstones covered what used to be a hearth in a depression. It was possible
to analyse a set of smithy wastes like PCB skulls, other iron slags and hammer scale
(Dunikowski et al. 1998, 145 - 152).


Thanks to voluminous work by Michel Mangin et al. (2000) the existence of numerous road
and village smithies are listed in the surroundings of Mont Auxois, ancient Alesia in Gaul,
especially in Haut Auxois, the south-eastern hinterland of the Gallo-Roman centre. There
are 55 forges mentioned in the monograph which were located in rural agglomerations
and villages and yielded cakes and amorphous pieces of iron slag and hammer-scale.
Forty samples from a set of sites were mineralogically and metallographically analysed by
148 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Philippe Fluzin (o.c., Chapter II, 165 - 268, Pl. XV - XXX). Taking Mont Auxois-Alesia
as a starting point we find more significant workshops within the range of about 15km,
e.g. Flavigny, Gissey, Darcey, Boux, Verrey, Villy, Blessey-Salmaise. Sombernon is more
distant (about 30km) as is Arcenant (more than 50km). The following sites provide a
more lucid picture as regards the looks and function of village and road smithies.
Gissey-sous-Flavigny (Hauteroche) was a rural settlement or villa with adjoint feature
called a monument. The northern annex (A) of the principal building housed a forge,
indicated by PCB and amorphous iron slag. Another smithy was located in the south,
across the Roman road 6. According to old excavations it was in partly preserved stone-
walled room or house (7m x 6.5m, destroyed in the west). The smithy in the eastern corner
of the structure worked in the final phase: a rectangular stone-lined hearth was placed
at the wall. Two metres from that was a small rectangular pit, presumably marking the
anvil position. The floor between these two features was covered with a layer of hammer
scale (Mangin et al. 2000, 134 - 136, Figs. 21 and 22). A lime-kiln, previously regarded
as pottery kiln, was uncovered in the same room.
The rural smithy in a small hamlet called La Montagne (07) situated at Vitteux and
Villy-en-Auxois was surrounded by a system of walls and was of a quite different character.
It consisted of four walled houses. In the central one (E 03) and adjacent to wall M 01
there was room H 5, a smithy (10m2 ) and probably a light shelter the floor of which was
covered with wastes: amorphous iron slag and a dozen PCB’s in an ashy black soil. The
space was not excavated and no hearth was observed (Mangin et al. 2000, 130, Fig. 20). It
resembles unexplored smithies at Blessey-Salmais, about 20km to the north-east. Blessey
is the most important site providing information about Gallo-Roman smithing activities
in the region discussed. It was located on the road (‘Chemin rouge’) leading from Alesia
to the Seine river springs where a sanctuary played an important role (about 1km away).
Blessey-Salmaise was a large rural agglomeration about 14km far from Alesia. The three
smithies H 301 (called ‘village forge’), H 407, and H 605 were not excavated but merely
cleaned and yielded some hammer scale and PCB slag cakes. In the westernmost part
of the habitation complex was a precinct with several stone-walled buildings. A large
sheltered forge F 104 (about 9m x 5m) was uncovered adjacent to the main structure
and its courtyard, leaning towards the eastern wall. Three hearths were installed in the
centre, one said to be constructed, perhaps elevated, the two other sunken. About 10m3
of ashy soil permeated with hammer scale and thousands of amorphous slag pieces and
rare PCB calottes were removed from this smithy which was in use over a long period of
time (the 1st to 3rd centuries AD) and produced mainly shoe-tacks (welcomed by pilgrims
and travellers from the Seine sanctuary). This is supported by Fluzin who investigated a
small rod made of pure ferritic iron (15mm long, 16mm x 11mm in section) which might
served as a blank; in addition seven different slag types were examined showing that
the reforging of iron pieces with heavy slag content had possibly also taken place in this
smithy which used different types of hearth. In sample 104/13 a droplet of cast iron was
mentioned (Fluzin in o.c. 201 - 210, Pl. XVII - XX; slags from the other, unexcavated
workshop show, according Fluzin, the work with unpure iron, recycling and refining which
were possibly carried out during different phases of the workshop’s life, see 210 - 216,
Pl. XX - XXII). At the crossing of Roman regional roads 13 (direction Autun) and 14
(direction Semur), existed the partially excavated (buildings I to V) wayside settlement
of Sombernon, which covered an area of 10 hectares. Building I, just at the crossway, had
smithy 01 (not described, a piece of hypereutectoid steel analysed), in building II forge
THE SMITHIES 149
08 displayed two hearths (100cm x 70cm and 120cm x 40cm, both 20cm deep) which
produced small smithing slag cakes, one concealing hammer scale traces and another one
some ferritic iron. In room 1 a small hearth was installed (60cm x 80cm); iron slags
and few bronzework waste are reported. Room 5 contained blacks soil and slags. The
workshop served travellers during the 1st century AD, although the habitation was in use
up to the 4th century (Mangin et al.. 2000, 127 - 128, Fig. 19; analyses by Fluzin 216 -
224, Pl. XXIII - XXIV).
Westwards from Nuits-Saint-Georges is the forest of Arcenant hiding ruins of a Gallic
fanum. During the second phase of its existence (3rd century AD) a forge was installed
at the wall of the northern courtyard (no 3) with an adjacent hearth, flagged with two
flat paving stones leaving the eastern side free for access (50cm x 50cm, 30cm high) The
feature was filled with earth on reddish subsoil and sealed with a yellow clay layer. A
depression is mentioned in the lower leve land two PCB slag cakes and a fragment with a
tuyere inlet. Slags and hammer scale were spread over the environment of the hearth as
well; (Fluzin in: Mangin et al. 2000, 258 - 260; slag analyses 260 - 263, Pl. XXIX). The
question whether the fanum still existed when the smithy was in operation remains open
(Mangin et al. 2002).

Smithies in Roman villae rusticae


These settlements represented a specific economical units the owners of which aimed at a
self-sustaining life. They harvested fields and orchards and tried to get necessary products
of various crafts by themselves making maximum use of slaves. In other words, apart of
some luxury goods they aimed to buy as little as possible and to be independent. The
blacksmith’s work was one of the very important activities and workshops have been
discovered in many of the explored villae rusticae.
A good example is the villa at Bibersist-Spitalhof near Solothurn (Switzerland), ancient
Salodurum in Germania Superior. The construction of a motorway prompted the exca-
vation of about 40% of the ancient built-up area in the rural NE corner of the complex.
Close to the enclosure a workshop was situated which functionned in two chronological
phases: I the late 2nd century AD, and II (3rd century AD, see Schuzany 1994). The first
smithy was built of wood (4.7m x 5.3m, with a lateral shed), flowed round by a brook
which separated this workshop from pebble, tile fragments and wall stone paved area with
traces of fire and containing iron slag and corroded small pieces of iron (gromps?); this
feature was explained as a reheating platform for working imported blooms. The smithy
proper had a central rectangular hearth (60cm x 80cm) built up of tiles and bordered
with vertically set tile parts (20cm high). It was fixed in a gravel-and-clay mixture. A
slag and charcoal layer covered the interior (23kg of slag per m3 , mostly in shape of PCB
calottes). In front of the hearth there were two rotary querns which probably served as
anvil supports. Finally, the smithy was flooded by waters of the brook and was later
reconstructed on the inundated terrain as workshop 2. Its ground plan was poorly pre-
served, marked by roof tile fragments (about 8m x 7m). The rectangular hearth (80cm x
100cm) was constructed principaly in the same manner of tiles and stones which formed
a kind of lining interrupted on the southern side where a channel was adapted, covered
with the imbrices, ridge tiles). This was an air-duct bearing to a recangular pit in the
virgin soil (90cm x 60cm, preserved depth 20cm). This was fed with falling water, which
came from the brook from west (but this section is merely postulated, not preserved) and
150 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
evacuated to the north by another channel back to the brook. In sum, the running and
falling water produced bubbles and pressure and the pit acted as a trompe inducing the
air into the hearth. Smithing iron slags similar to those from the first smithy covered the
indundation level (ibidem, 144, Figs 14 to 20, here Fig. 59: 1 - 2).
The following two cases have been discussed above (hearths, see Chapter VIII) so
they be only briefly described here. The two groups of large circular hearths (6 and 10
features) at Riom, Switzerland, do not offer any sign of a regular smithy, despite the
fact that working of iron was carried out on both spots (slags, hammer scale, tuyeres).
Therefore, the opinion expressed by Rageth (1982, 204, 205) that these installations were
used in course of extensive building operations in this settlement in the 1st century AD
can be accepted.
The farm and estate at Horath near Trier (W. Germany) yielded an unroofed room 7a
(5m x 3.5m) with the discussed domed hearth and PCB slags (Cüppers 1967, 129 - 131,
Fig. 10). A battery of 5 semi-circular burnt-red hearths (diameter 50cm, depth 20cm,
with inclined air-inlet 20cm x 20cm) and 48kg of iron slag (larger and smaller PCB’s,
0.5 - 1.5kg and 0.04 - 0.2kg) revealed the re-excavation (1992) in a courtyard of the villa
at Jemelle, Namur Province in Belgium (Mathieu et al. 1994, 141). Clear signs of the
smithy structure are lacking. The authors take into question that the hearths were used
both in forging and reheating operations.
Attention should be paid to a villa rustica in Brittany - Pont-Croix-Kervenennec where
a squarish stone-flagged hearth (100cm x 100cm) dominated the centre of a walled room
(4.5m x 4m). Iron nails were found in a water tank at the eastern wall which would
have been fed with water through lead pipes from the roof. Some crucible fragments were
found in a room corner but no slag is mentioned (Galliou 1984, 80 - 81, fig. 7, here Fig.
57: 5).
Several smithies were reported from estates in the Gallia Narbonensis (the south of
France). On the left bank of the Rhône river, NW from Avignon, a villa was investigated
near Ramière, dating from the late antiquity (the 4th century AD). A smithy was installed
in the eastern end of the central building (20m x 10m) consisting of an oval sunken hearth
(78cm x 98cm) with a possible bellows bed on one side. There was a limestone block of a
trapezoidal shape (with marks of forging) - an anvil, surrounded with a layer containing
hammer scale. A small stone wall was set behind the hearth and the anvil from the east.
Another small hearth occured at the northern wall. There was a pit filled with charcoal
reserve (box-tree, oak) in the SE corner. Iron slags were dispersed over the floor of the
workshop, especially in the southern part of the room. In addition, outside the building
at the southern wall there was an adjacent domestic kiln or oven constructed of tiles. Iron
objects were found in the room (a knife, a bent blade, a fragment). The function of the
smithy consisted rather of the maintenance of iron artefacts than in their manufacture
(Maufras and Fabre 1998, 216, Figs 3 to 10).
Not very distant is Mayran (Saint-Victor-de-la-Coste, Gard); a villa existed there in
the late antiquity. Southwards of a stone building two accumulations of iron slag appeared
(1220 and 1001, 63kg excavated, mostly PCB’s). They attest to the activity of a smithy
which was not revealed in the area of excavations (Buffat and Petitot 1998, 175 - 180,
Figs 2 and 3).
Further south three sites with indications of smithies were found in Hérault. At Domer-
gue à Sauvian (Ginouvez et al. 1998, Figs 1, 2, and 3) a villa from the 4th century AD
housed what is described as two elongated hearths 331 and 339 (unexcavated) in room
THE SMITHIES 151
F. No slag is mentioned but an iron block or bar weighing 1920g which corresponds with
5.68 Roman pounds, maybe a divided part of a larger block is reported.
The villa rustica at Prés Bas (westwards from Sète) prepares some difficulties in inter-
pretation of the blacksmith’s activities. Iron slag in shape of calottes was found in the
area having been dispersed during the life of the estate (the 1st, late 2nd and the 4th
centuries AD). However, it was in the 4th century when on the periphery of the residental
building that a wine cellar (300m2 , with 90 dolia) was abandoned and a workshop tem-
porarily installed inside. An oval hearth built of tiles (a large one, 10cm x 120cm) adhered
to a burnt-red spot. A limestone block served apparently as an anvil but unfortunately
no slag was found around: instead of that about 400 nails and agricultural implements,
some of them constructed by welding came to light. The workshop did not produced any
artefacts but was destined for repairing the domestic iron (Pellecuer 1998, see Figs 4 to
6, and 9).
At Saint-Jean d’Aureilhan, periphery of Béziers, a villa (the 1st to 6th centuries AD)
yielded an oblong hearth the terminal part of which was well fired and led into by a
bellows bed. The installation was surrounded with iron slags and ashes. It is interpreted
as a small smithy for occasional work. Date: around AD 400 (Ginouvez and Vidal 1998,
Fig. 3).
Two smithies were revealed in the coastal region of Catalonia. The villa at El Vilarenc
(Fig. 57: 4) had a rectangular building block in the southern part of which several rooms
included a basin, a hypocaustum, and a forge (4m x 7m). The latter was equipped with
two hearths, one of which was big and circular, and a rectangular one in the corner at
the western wall. These installations were treated in Chapter VIII. Slags, ash, charcoal,
iron were found there. As to the dating, the 1st century AD comes into question, just
after the founding of the Roman province Tarraconensis. Later was the villa at Tossal
del Moro (the 2nd - 4th centuries AD), north-eastwards. A valled structure (6.2m x 4m)
housed a central stone hearth, probably elevated. Iron slags and iron keys were found in
the fill (both sites se Pérez-Suñé it et al. 1998, 223 - 225, Figs 2, 3, and 5).


Turning eastwards a 4th century AD villa was investigated at Petöháza in W Hungary in
the province Pannonia Superior and is worth of mentioning (Gömöri 2000, 132 - 136). A
freestanding post hut 4.2 x 4.2m south of the stone-built complex functioned as a smithy.
The sunken oval hearth (about 30cm x 40cm) with a stone protection of bellows was near
a larger pit. Beside of iron slags a part of a hinged smith’s tongs and some iron scrap
came to light (o.c., figs 83 to 85).

Iron Age smithies of Central and Northern Europe


In times which saw the early development of the fully-fledged technical Iron Age in Greek
and Roman worlds, northern Europe lived in the Bronze Age (Montelius’ periods III -
VI) and central Europe witnessed the coming of the Early Iron Age (in archaeological
terms). Few early smithies were recorded in these regions. It would be logical to find the
earliest of them in geographical tracts neighbouring the advanced civilizations. However,
radiocarbon data have enflamed a discussion cencerning three sites far in the North, in
central Sweden. The first of these is Linga where a smelting bowl furnace and two possible
152 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
smithy hearths yielding heavily wüstitic slag and pearlitic iron particles the date of which
was presented as 1377 BC to 212 AD, 965 BC to 460 BC, and 573 BC/357 BC to AD
410/640. A slag dump of 2.3m3 is mentioned. Another site, Hallunda, yielded 450g of
iron slag (apart of copper smelting waste), dated to 1187 BC, 1048 - 850 BC, 889 BC to
598 BC, and 828 BC to 452 BC. Hällby is the site where 5 stone-lined hearths/furnaces
(internal dia. 50cm) came to light: A 14, A 15, A 20, A 28, A 52, and an anvil stone. Plus
a part of a possible iron bloom (dia. 7cm). Radiocarbon chronology: total range 2150 -
2201 BC to AD 596, i.e. 1213 BC to 930 BC, 1004 BC to 827 BC for charcoal in slag,
760 BC to 380 BC for furnace A 52 AD 260 to AD 596 for charcoal (Hjärthner-Holdar
1993, 61 - 73, 74 - 76, 80 - 86, Figs 50 to 52, and 69 to 73). Should be the earliest data
taken into account, we face to a curious anomaly in terms of cultural and technological
diffusion. Hjärthner-Holdar in her praisworthy work (o.c., 26 - 33) attributes importance
on certain influences from the Ananino culture in the Kama river region, northern Russia,
especially what concerns some Swedish bronze castings, and seeks to promote the idea of
inseminating elements of ironmaking from that direction as well. However, in the Ananino
culture (the 8th to the 3th centuries BC) the bronze technology was predominant still
about 500 BC and was replaced by iron not earlier than during the 4th - 3rd centuries
(Koryakova 2001; Kuzminich 200l). Ananino has to be held as retarded in terms of the
spread of iron technology. Should be the later dates for the Nordic Bronze Age accepted
(as Chr. Zimmermann proposed in 1998), the chronology of the workshops in question
would fit with the Late Bronze Age and Early Iron Age in the Continent.
No serious dating problems face other early smithies in central and western Europe,
according to the Hallstatt D/La Tène A pottery they date from about 500 BC. Instead
of that other problems arise. The cave Býčı́ skála (Bull’s Rock) in the Moravian Karst is
known by a monstrous princely burial or sacrifice place, discovered in early 1880’s (Wankel
1982, 266 sq.) Since that time the site has steadily been the subject of discussions as to
what actual meaning have had the massacre of men, cut-off arms, chariot remains, bronze
and pottery vessels and burnt cereals had. In the eastern cave wall a ‘smithy’ was placed,
yielding 4 heavy iron sledge hammers (about 8kg each), 2 iron block anvils, 1 chisel, 1
pincer tongs an a half of a double-pointed iron bar of the early type, slag, hammer scale
and casting moulds (Collections of the Naturhistorisches Museum in Wien; see Ohlhaver
1939, 145, PL. 7; Pleiner 1958, 86 - 89, Fig. 13; idem 1962, 60 - 62, Fig. 10; Souchopová
1996). It is also not clear whether the smithy existed before or during the hecatomb. The
place was separate from the bulk of sacrifice finds, nevertheless it was covered with ash,
charred wood or charcoal (as to J. Wankel, the discoverer).
The Waschenberg site (near Bad Wimsbach, Austria) is remarkable in two respects:
it was a fortified locality protected with a moat system and it contained and area of
ironmaking (bowl furnaces). Ironworking was also attested during the rescue digs caused
by gravel expoitation. Three features were labelled as smithies: smithy S (a shelter 3.2m
x 3.8m) represented a combination of a bloomery bowl furnace and a smithing/reheating
hearth (dia 60cm, 30cm deep), with an adjacent stone anvil with stroke traces. Smithy R
(2.6m x 3.1m, sheltered with a post structure) involved a hearth (dia. 50cm) containing
smithing slag. A flat stone-lined hearth was situated in front of a post-built shelter which
served as smithy M (3.2m x 2.8m, marked by post holes). Iron smithing slag was present
round about (Pertlwieser 1970, 58 - 62, Figs. 6 and 7). The treatment of iron blooms or
sponges must be assumed. Any manufacture of artefacts cannot be proved, although it
comes into question as well. No non-ferrous metallurgy was reported from this context.
THE SMITHIES 153
Rescue excavations at Celle-sur-Loire, site of ‘Bois du Jarrier 3’ produced another work-
ing place where iron also was heated and forged in connection with reheating operations
(see Chapter VIII). No smithy ground plan could be made out but two the remains of
two hearths emerged in the subsoil. ‘Foyer l’ (60cm x 50cm, 15cm deep, with an air-inlet
slagged clay fragment in the debris of slag and charcoal) and ‘Foyer 2’ (85cm x 70cm,
15cm deep) with traces of burnt-red clay lining. The third hearth (‘Foyer 5) was sev-
erly eroded before the excavation. Slag concentrations (96kg in total) were associated
with with the above hearths: PCB cakes (53 kg) and amorphous pieces, both lamellar
and globular hammer scale (6kg, from that 2kg at hearth 2) were identified by P. Fluzin
who analysed the wastes. What is remarkable is that a furnace bottom (diameter 24cm,
weiging 10kg) was found which indicates that reheating of blooms took place on spot.
Particles of highly carburized steel were registered including a droplet of white cast iron
(ledeburite, more than 1.7% C). The site is situated not far from Puyssaye, later the
renowned iron producing region. However, the site of ‘Bois du Jarrier’ was dated to much
earlier times, some pottery fragments are of the Hallstatt/La Tène periods and radiocar-
bon measuremnts show the time span of 761 BC to 412 BC. The production volume was
declared as small-scale (Orengo et al. 2000, 45 - 66).
Hallstatt/La Tène pottery and Attic sherds characterized the recently discovered work-
place at Sévaz-Tudinges, near Lake Neuchâtel, Switzerland. Again, no exact ground plan
could be found. Three pits (diameter about 180cm) filled with slag (more than 140kg,
stray finds in the environment included PCB’s, hammer scale 9kg), a charcoal combustion
feature and 18 rusty iron fragments were registered. V. Serneels classified the production
at Sévaz as small-scale volume; it included bronzeworking as well which is demonstrated
by several crucible sherds (Serneels and Mauvilly 200l).


This survey has to be completed by a reamrk about the Scythian fortified site Belskoye
(ancient Gelona) in the Ukraine (the 7th - 6th centuries BC, Shramko 1987). In area 26
(1975/76) a hearth (dia. 65cm, with burnt-red walls) was excavated as well as quartzite
and granite anvils (81.5cm x 32cm x 16.5cm), a whetstone, and an iron chisel. About 40
other iron artefacts and smithing slag are mentioned in the description.

Celtic and Dacian smithies

Celtic oppida were centres of political and economical life of considerable social entities.
Manufacturing craft activity was an integral part of their function. Excavation revealed
an enormous number of iron artefacts (about a hundred kinds from fittings to perfect
tools) so that smiths must have worked in these agglomerations. In several oppida smithy
workshops were actually found or traced. Before treating we have to deal with three
localities outside these centres. All of them are connected with the primary production
of iron - with bloomery ironworks.


The site at Kundl-Lus, lower Inn river valley, Tirol, which was submitted to a rescue dig
in 1984. Within the unearthed triangular area several smithing slag concentrations were
154 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
observed which also involved bloom fragments, iron bands and rods and sheet pieces. The
slags (incl. the PCB calottes) were wüstitic and fayalitic, permeated with lamellar and
globular magnetic hammer scale. The distribution of bloom parts and iron objects indi-
cate a reheating area in the south (‘Grobschmiedearbeitsbereich’) and the area of further
ironworking in the east (‘Feinschmiedearbeitsbereich’). The iron in Kundl is poor in phos-
phorus (0.036% in average) but rich in sulphur (0.058%) and nitrogen (0.392%) which
unusual for bloomery iron (a completely different in comparison with phosphoric iron
from Manching, for example). The composition was explained by the use of cow dung in
the heating and smithing operations (Maurer 1993, 323). No hearths are mentioned.The
date of the site is La Tène period C.
Then, there is Mšec in western central Bohemia where a bloomery was investigated in
the 1960’s (Pleiner and Princ 1984). In course of the second half of the 2nd century BC
several phases of iron smelting leaving a slag-pit furnace cluster. The first step of the
reheating of iron sponges produced was performed in a horseshoe-shaped hearth where
the metal was separated from the residues of bloomery slag blocks. Further refining
followed in a smithy (Fig. 56: 5). This was a sunken-floored feature two meters away
from the slag-pit furnaces and the reheating hearth; it measured 4.4m x 3.5m by 30cm
depth. There was a stone anvil (46cm x 35cm x 16cm) surrounded by hammer scale
collected by a magnet, in the centre. In the southern part of the hut (which must have
been roofed) there was a small elevated platform covered with PCB slags of about 10cm
dia., hammer scale, and charcoal (spruce). A hearth must have been placed there but the
fire pit had been apparently destroyed and removed. The smithing slag was not analysed,
in contrast with the bloomery waste (o.c., 148 - 151, Fig 6 and 11: 10). With exception
of a corroded small iron bar no iron objects or tools appeared in the fill. The chronology
of the complex was synchronous with the floruit of the Bohemian oppida. The nearest of
them was Stradonice, some 20km south-westwards.
A different situation appeared on the periphery of the Celtic cultural province, in the
Siegerland of W Germany, known as a La Tène ironmaking region. Laumann (1985)
described a so-called ‘Schmiedepodium’, a smithy terrace-like platform near Neunkirchen-
Zeppenfeld. Unfortunately, no smithy ground plan could be traced but a fire-place (100cm
in dia.) yielded PCB slags and a stone anvil was also found. No post holes marked
any trace of a shelter. Close by, in a dwelling area (‘Wohnpodium’) a forge spoon was
found.The smithy in the Minepit Wood area, Sussex, England (Money 1974) must have
been connected with iron smelting. The hearth has been described in Chapter VIII. No
smithy in the usual sense could be distinguished.


In the 1st century BC the oppidum or, rather, the town on the hill of Magdalensberg
in Carinthia was a leading centre producing iron ingots including artefacts called anuli
(rings), unci (hooks, clamps), secures (axeheads) and incudes (anvils), which were made in
considerable quantities. The transactions are inscribed on the walls of tabernae eastwards
from the later temple area (Egger 1964, 28 - 29, 30 - 32) and the iron artefacts were
distributed in sets of hundreds of pieces to Italy and as far as Africa; it is not excluded
that the matter was of special kind of stylized ingots and bars of the famous Norician
steel here. Certain smithy installations were already mentioned in Chapter VIII and in
the paragraph dealing with urban smithies. Briefly, at Magdalensberg 11 sites (mostly
THE SMITHIES 155
tabernae) were reported where smithing activities were performed, especially around the
later forum. All of them were damaged by subsequent rebuilding so that no complex
reconstructed (Dolenz 1998, 15 - 45).
Turning north-west- amd westwards to Gaul we come across Rheinau, an oppidum
close to the Rhine river falls, near Schaffhausen in Switzerland. Three pits denoted as
Schmiedeessen (smithing hearths) and PCB slags, hammer scale and brick-shaped tuyeres
were reported after a rescue operation in 1994 (Schreyer and Graf 1995). About 30
brooches and small iron artefacts of iron and bronze and 14 silver coins supplemented the
finds assemblage.
In many of the European Celtic oppida the smithies worked near the fortification
gates. This also was the case of Bibracte in Gaul (Mont Beuvray). A smithy was already
described by Bulliot (1899 I, 2 - 7) during the early excavations in the 1890’s. It was a
sunken-floored room (6.5m x 5.5m) situated outside the rampart system but close to the
fortification walls. The floor was paved with sandstones where a print was left marking
the position of the sunken base of an anvil. Blacksmiths’ activity also concentrated at the
hollow way called Come Chaudron (ibidem, 76 - 79; Thiollier 1899). Recently a smithy
has been reported from the area at the Reboul gate. It was a rectangular hut (4.5m
x 4.5m) framed on all sides by post gullies. A dozen small hearths (diameter 20cm to
50cm), pits (anvil bases) and the bottom of an amhora (water tank?) are reported to
have been found inside. Hammer scale waste is also mentioned. Among the finds a file
and a small hammer, iron wire, semi-finished brooches, sheet and copper traces deserve
a mention (Fichtl 2000). Fig. 56: 4.
From central Gaul there is reported a smithy excavated at Clermond Ferrand (Le
Pâtoral), unfortunately not illustrated by a plan; just 180kg of smithing slag (PCB slags,
both complete and fragmentary), globular and lamellar hammer scale and pieces of block
clay tuyeres, dating from the 2nd century BC are mentioned (Orengo et al. 2000, 123 -
124). The smithy at Feurs (Forum Segusiavorum), Maison de la Commune dates from
a later period (1st century AD). The incomplete plan of the workshop (not reproduced)
involved two hearths, one built of tiles, the other sunk in the floor (45cm x 70cm, 15cm
deep), full of hammer scale and surrounded with smithing slag (PCB calottes, 5kg in
total) and block tuyere fragments (o. c. 122, fig. 3).
As regards workshop ground plans, the excavators of Manching, an important oppidum
in Bavaria, were less lucky although blacksmith’s slags and tools came to light in con-
siderable numbers spread over all the excavated areas. Among the slags there are PCB
cakes (Jacobi 1974, 246 - 247, pl. 105: 1 and 2; examined by Keesmann and Hilgart
1992). Accummulations of this waste were also observed in the site of the ‘Altenfeld’ (pit
798c with 42 pieces, pit 1230a, ditch 1003-1 with more than 7kg; ditch 1523-1).About
100kg of slags are mentioned in total, and numerous blacksmith’s tools like hammers,
files, forge spoons and artefacts of every kind were found (Jacobi 1974, passim; Sievers
1992 and eadem in Vorbericht Manching 1996-1997, Germania 76/2 1998, 642, Fig. 8;
eadem 1999, 20). A smith’s hoard in pit 822b4 from the latest phase, La Tène period D1,
deserves special attention. It was originally wrapped in a cloth and comprised iron tools
and bronze objects and numerous pieces of iron semi-finished artefacts and scrap (Sievers
1992, 201, figs 96 and 97, pl. 57 and 58). Elongated iron bars and their fragments, rods,
wire, fragments of bent and overlapped sheet attest that the blacksmith’s production at
Manching must have been very intensive during all chronological phases and consisted of
the forging of artefacts from bars and recycled scrap metal as well as the reheating and
156 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
refining of imported blooms.
In central Europe three Bohemian oppida have to be named. Závist near Prague,
where long term excavations took place, had a smithy behind gate D, marked by many
PCB slag cakes, brick-shaped clay tuyeres and the like. More important was the smithy
of an estate in the acropolis where gradually a workshop was identified and reconstructed
after the evaluation of several seasons of investigation could have been commented (Fig.
56: 6). This was a space 8m x 11m with a shelter marked by post holes (7m x 3.5m).
In its centre a large oval hearth left traces and there was a stone anvil close-by. Another
elongated hearth was outside the shelter in the open air. PCB calottes were collected
(6kg) but a very large innumerable amount of waste was thrown down the slope of the
inhabited platform. Finds: 2 anvil stakes, part of hinged tongs, 2 forge spoons and scrap
iron. A squarish feature close to the anvil stone is of not known function; the excava-
tors do not rule out a kind of altar (Drda 2000, 18 - 21; the exact description should be
published: Drda, in press). At the oppidum of Hrazany, south of Prague on the Vltava
river, the blacksmith’s work was performed at gates A and B. No ground plan of any
smithy stands is available but a lot of PCB slags were present and a clay block tuyere,
a hearth spoon and a hammer are mentioned in the debris (Jansová 1960, 674, Figs 240:
2 and 7; eadem 1999, 172). In south-eastern Bohemia an oppidum known as Hradiště
u Lhotic was excavated during several seasons in the 1970’s. In course of these inves-
tigations a free-standing post hut was recognized (3.8m x 3.3m). In the central part a
pit with a stone was found: the presence of hammer scale supports the opinion that this
was an anvil. In total, about 120 iron fragments, a forge spoon, broken iron brooches,
burnt-red clay lining pieces and charcoal, as well as crucible fragments were uncovered
which indicates that a smith producing small iron ornaments and mastering the bronze-
work was active at that place. Unfortunately no hearth in situ was identified (Princ 1981).


In the cultural province of the Dacians on the territory of modern Romania the most
famous is the capital Sarmizegetusa (Gradiştea Muncelului). In the literature, smithing
operations are mentioned on terraces where rectangular houses with stone sustaining walls
were excavated but not described in detail. Elongated hearths and several hoards of iron
objects are located on terrace VIII (Daicoviciu et al. 1953, pl. II). Otherwise Sarmizege-
tusa is known for finds of blacksmith’s tools (block anvils, 4 tongs, hammers, files, wire
drawing dies (see Mărgitan and Andrit.oiu 1971; Glodariu 1980). The situations are pre-
Roman, dating from the late 1st century AD.

Germanic rural smithies

From the beginnings of European ethnology territories north of the Danube and east
from Rhine to the Germanic branch of people. Since the Iron Age was infiltrated to these
parts of central and northern Europe, iron was made and worked within this geographical
tract and, naturally, in certain places smithies have been identified. The earliest of them
were revealed in Scandinavian lands, in Denmark and Sweden (about the 1st century
BC). Hodde, W Jutland: a rural fenced settlement with freestanding longhouses com-
bined with cow sheds is dated in its earliest phase to the late pre-Roman Iron Age. In the
NW corner of the palisade a smithy workplace (about 5m x 5m) was adjacent to one of
THE SMITHIES 157
the longhouses. Slags and hammer-scale were reported. Later, in the Romano-Barbarian
period (the 3rd phase of the inhabitation) another smithy appeared in the south of the
palisade but outside its circumference. Both served apparently as repair workshops main-
taining the iron inventory of the inhabitants (Hvass 1975, 142 - 158, fig. 10; Müller-Wille
1983, 228). In the same region Becker (1980, 59 - 60) investigated an isolated complex
at Grønbjerg Skole where a minor feature and numerous iron slags occured close to two
longhouses which indicate the existence of a smithy (also see Müller-Wille, o.c., 228).
Becker also reports of a small post-built feature (4m x 3m) revealed at Omgård (1982, 67
- 68, fig. 17), where slag and hammer-scale were observed. It was possibly a smithy. Two
hearths with stone-lined rims (dia. about 35cm) but no smithy complex were reported by
Arbutsro, Södermanland, Sweden (Hjärthner-Holdar et al. 1996.


Reports on early Romano-Barbarian rural smithies are relatively scarce. Recently, a small
sunken-floored hut F (2m x 2.8m by 5 - 10cm depth) was published from Beroun, SW
from Prague (Jančo 2000). This was interpreted as bloomery with residues of a shaft
furnace (diameter 35cm x 40cm) in the SE corner, and with an anvil stone found nearby
sticking in the floor. The Beroun site yielded, in fact, iron smelting evidence but the
apparatus were slag-pit furnaces (unpublished) of an entirely different character. Hut F,
dated to A/B2b periods (1st century AD) should most likely be interpreted as a smithy.
The hearth did not contained any iron slag; various pieces were dispersed in all of the
neighbouring pits but hut F also had fragments of bronze sheet from vessels in its filling.
Unfortunately, the feature was severely eroded and only the lowest parts were preserved.
A very poorly documented sunken feature with several bays came from Kapalica near
Jarocin, Poland. A stone (anvil) was in one of the bays, a stone-paved fireplace in the
centre and the remains of a furnace/hearth in the northern bay. Rauhut (1957, 38 - 39,
Fig. 33) is inclined to see a smithy.
Smithing operations related to bloom reheating and reforging must have taken place
at Pokrzywnica near Pawlów, in the Holy Cross Mountains iron production centre in
central Poland (Bielenin etal. 1996, 356 - 365, Figs 12 to 17). Three hearths (F 11, F
12, and F 27) were investigated. F 11 was built of bricks and equipped with an inclined
air-inlet (diameter 95cm, depth 45cm), hearth 12 (115cm x 85cm) was sunken in a pit
and contained 40kg of PCB slag. Hearth 27 (120cm x 100cm) yielded 26kg of PCB slags
which occured, moreover, in the eastern part of the complex investigated. The date: AD
150 - 200.
Finally it is unavoidable to mention anew the Huckhoe site in Northumberland, Eng-
land, with the rock-cut hearth (see Subchapter 6) and its enigmatic dating: originally
2nd century AD (Tylecote 1962, 232 - 233), subsequently, according to later added radio-
carbon measurements, redated to the 7th - 6th centuries BC. However, the hearth is not
mentioned by Jobey in his report on radiocarbon measurements of the palisade Bronze
settlement of Huckhoe.
In terms of investigation of smithies dating from the late Roman-Barbarian or Migra-
tion periods (the 4th to 6th centuries AD) an important aspect has to be stressed: most
of those we know of nowadays, were in connection with bloomery ironworks indicating
that their function should be bound to that activity - that is the reheating or reforging
of the blooms produced.
158 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
At Süderschmedeby, site Auberg (Schleswig-Hollstein, N Germany), Hingst (1978, 67 -
69, Fig. 80) announced a ‘roofed smithy’ near a slag heap having been left after ironmak-
ing activities. A large anvil stone (80 cm) wedged by stones at its bottom was surrounded
by slags in shape of PCB skulls. Four smithies were excavated at a well known smelting
site with slag-pit furnace clusters (about 500 units in total) at Joldelund, site Kammberg
in the same region (the late 4th - early 5th centuries AD). It has been noted that the iron
production and the rural inhabitation came to pass in several chronological and strati-
graphical miniphases so that the furnaces, longhouses and individual smithies appeared
as non synchronous. The excavation was realized in test trenches and three main areas:
the north-western, the south-eastern, and the southern, each subdivited into sections la-
belled with letters. In area G (north-west) the smithy represented a post-built shelter
(house 12, 6m x 7m) which included a central debris depression (200cm x 130cm) and two
non-synchronous hearths (100cm x 60cm, and 60cm x 40cm, the latter being younger).
Several spots inside containing hammer scale signalize the forging of heated iron on an
assumed stone anvil which was not left in situ. In the north a large deposit contained
metallurgical waste, incl. some PCB slags, having been analysed by Ganzelewski. This
deposit seems to break slightly the smithy ground plan and could be a little later. The
south-eastern excavation brought in area M another feature explained as smithy (post-
built house 14, 10.5m x 5.5m) but no hearth could be identitied, just a waste depression
(75cm x 150cm). In the south a large debris deposit extended, which also yielded, among
others, some PCB calottes (analysed). The chronological relation is not very clear, since
the two post holes in the southern house wall, marking the entrance, lead directly to
another debris concentration which adjoined in distance of some five metres. About 7
metres eastwards another smithy was uncovered (excavation area T, house 15) which was
preserved as a ground plan marked by posts only in its eastern part (estimated dimen-
sions 4.5m x 6 - 9m). A debris pit and a twin hearth were situated in the possible centre.
In the south a slag and waste heap adjoined and might have preceded the smithy hut.
The southern excavation comprised test trenches and was rather limited since a wooded
terrain did not allowed a large-scale dig. Under a layer containing slag and charcoal the
SW corner of a post building (house 20 in area R) could be recognized which, neverthe-
less, showed the existence of a debris pit (90cm x 100cm) damaged by three later slag-pit
furnace remains. There was a pit containing stone fragments close by, possibly the anvil
position. A smithy was evidently destroyed there after the smelting operations have been
renewed. The analyses of smithing slag (including PCB’s) and lamellar and globular
‘Schweissperlen’ hammer scale brought no unambiguous results in terms of reheating and
refining the blooms or iron sponges or bars or forging artefacts. All of these operations
come into question and the smithies could have produced iron utensils and tools to inhab-
itants in certain ‘miniphases’ or work the iron smelted in other time sections. Fragments
of clay protection shields were noted in all of the waste deposits and were apparently inte-
gral parts of any hearth - a better preserved shield or ‘Essestein’ published Nikulka (2000,
74 - 75, fig. 10) from Heek-Nienborg, in the same region. The Joldelund-Kammberg site,
investigated as a VW-Stiftung project, was published in two monograph volumes. The
forges are treated in volume 1 (Jöns 1997, 134 - 142, figs 81 to 87); analytical data of
smithing slags are presented in volume 2 (see Ganselewski 2000, 63 - 64, 81 - 86).
The problem of the connection of smithies with an extensive iron smelting may be
illustrated on finds at the Danish site of Snorup, Jutland. More than 1000 slag-pit furnaces
were uncovered in clusters or rows during systematical excavation but only one feature
THE SMITHIES 159
represents a smithy: that on the southern rim of the densest (northern) slag-pit furnace
cluster between three longhouses which apparently belonged to different miniphases of the
site development - a situation quite analogous to Joldelund but, according to radiocarbon
dates, a wider chronological range is being given (AD 330 to AD 570). Unfortunately,
the very deep ploughing destroyed the floor level so that merely the most sunken features
could be registered: 6 post holes of a rectangular ground plan (4m x 3m) indicate a
roofed shelter (Fig. 58: 3). In the centre was what remained from the bottom of an
oval burnt-red hearth (75cm x 30cm, 13cm of the preserved depth) filled with black
soil, magnetic dust (hammer scale) and minute slag pieces (about 3cm). An important
detail is that a PCB slag cake (12cm x 15cm, weighing 1kg) remained on the top. Some
centimetres eastwards from the hearth a shallow depression (ca 40cm x 50cm) with dark
soil and hammer scale fill was situated, also containing fragments of slagged clay lining
and some small stones (10cm - 15cm), possibly residues of fixing a stone anvil which was
not preserved (Voss 1993, 100, 102, fig. 10; Voss, personal communication 2001). In fact,
from the archaeological point of view it is difficult to ascribe this smithy, a single feature
among hundreds of slag-pit furnaces to having worked for the benefit of inhabitants or to
having been part of a chain phase operation - the refining of blooms, although Buchwald
(2002, 10, 42, fig. 18) is ready to interpret the slag cake from the hearth as refining slag
(like the PCB calottes in general).

Figure 59: Smithies in the Roman villae rusticae and at the Romano-Barbarian Snorup. 1
- 2 Biberist-Spitalhof, Switzerland (1 earlier workshop, 2nd century AD, 2 later workshop
with a trompe air-duct system; 3 Snorup, Jutland, with a PCB slag in situ in the hearth.
1 - 2 after Schuzany, 3 after Voss.

Longhouses, slag-pit furnaces, smithies: again, these features characterized the long
term development of an estate-settlement at Vorbasse, central Jutland. The localization
160 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
of fenced farms changed during the 1st to 4th/5th centuries AD. Three ‘smithies’ are
mentioned as parts of longhouses in the 4th century AD phase, two in the western tract
of estates, and one in the central tract. However, they are indicated merely on the basis of
the presence of iron slags or of ‘small pits with slag’ (Hvass 1980, 161 - 180; Müller-Wille
1983, figs 6: 1 - 5, fig. 7: 1). Smithing was performed in this settlement in the Middle
Ages as well (see below). Larsson and Hjärthner-Holdar (2000, 13 - 15) write on a smithy
hearth, PCB slag, hammer scale and iron samples from Pornullbacken, Finland; no details
concerning a smithy complex are available.

Smithies in early medieval pre-urban centres


In lands which were not touched by the Roman administration and provincial system the
economical progress and social differentiation proceeded in a similar way as among Celts
many centuries earlier. Many fortified settlement agglomerations wit a rural hinterland
appeared in non-Roman regions: among Germanic tribes, Vikings and early Germans,
and among Slavs who have colonized a large part of Europe. These centres were econo-
mical and political focuses where crafts played an important role supplying not only their
own inhabitants but, at least partly, the rural environment as well. The blacksmith’s
work was an integral part of their life and smithies must have existed in all of them.
Under favourable circumstances, also smithies could be investigated during excavations
carried out in these settlement complexes. Following this discussion of smithing among
the Germanic people we turn to to the North of Europe and see that economic and social
stratification also developed there in pre-urban times during the Viking Age. Vikings were
not only sea-raiders. Their craft working attained an admirable level and many artefacts
were spread by trade.
When discussing blacksmiths’ activity, one site is of particular importance although no
smithy ground plan could be discovered during extensive excavations. It is Haithabu or
Hedeby which crowns the vik of the Schlei bay at Schleswig (9th to 10th/121th centuries
AD). In terms of matter that concern us here the importance is inherent in the wastes: the
smithings slag accumulations in certain parts of the semi-circular fortified settlement area
and in the harbour. Two were in the NW, close to the rempart and gates, amounting to
quintals in weight (area H and area D). Another place where smithing slag was registered
is the south, area G, adhering the rampart and the shore. The bulk of slags appeared as
PCB cakes, 5cm to 12cm in dia. which were carefully treated by Petra Westphalen who
devoted a monograph to the Haithabu slags (Westphalen 1989). 34 tonnes were yielded
and evaluated (o. c., 24). Among the material studied numerous clay fragments with
air-inlet holes came to light which signalize the frequent use of clay bellows protection
(as shields). What iron was worked at Haithabu? Probably, the production depended on
imported bars from Scandinavia; the PCB slags attest, with a high degree of probability
that no (imported) blooms were refined in the smithies of Haithabu but commercional
iron was transformed into artefacts. By the way, this was the conclusion made on a
Haithabu PCB in early days of smithing slag investigations (Pleiner et al., 1971). A similar
result was expressed by Thomsen who examined four other samples of slag calottes from
Haithabu (Thomsen 1971a, 103 - 109, Figs 5 and 6). It should be noted that Thomsen
puts the mark of equality between the German term ‘Ausheizschlacke’ (reheating slag)
and ‘Schmiedeschlacke’ (smithing slag); whereas his conclusion concerns smithing slags
alone.
THE SMITHIES 161
A more modest but not neglectable role played the port of Menzlin on the German
Baltic coast, a Scandinavian/Slavic trading post of the 9th - 10th centuries AD, with a
hinterland settled by Slavs (Schoknecht 1977, 64 - 65, fig. 21; 73 - 78, Pl. 39 and 40:
12). A smithy is assumed to have worked in slightly sunken house IV (4.6m x 9.8m). No
hearth could be identified any more but PCB slag cakes (8cm to 12cm dia.) of about
100kg in total and some tuyere fragments and iron bars were found in the fill. A stone
(anvil) was situated at the western long wall. Some slag calottes were chemically analysed
by F.-J. Ernst (ibidem, 152 - 155) who ascribes them to the ironworking. No blooms were
found at Menzlin; instead, various bar-shaped objects. A faggoted, folded and welded
piece piled from 4 iron bands is interesting (pit V/15, Schoknecht (o.c., 75, Pl. 40: 12)
as it suggests scrap recycling took place on the site.
Another type of settlement or residence from which ironworking evidence was an-
nounced is the ring-shaped Fyrkat in Jutland, Denmark, 10th century AD). Inside of
the ring-fenced interior three of four quadrants were occupied by square-arranged ship-
shaped longhouses with a central courtyard. Iron slag, stone bellows shields, crucibles and
moulds were found in all of them. The iron working waste being was most frequent in the
SE quadrant - eastern longhouse of the squarish formation (Roesdahl 1977 I,45, fig. 44;
Müller-Wille 1983, 243, 246 - 247, fig. 14). Iron and non-ferrous objects were made and
repaired on behalf of the inhabitants in Fyrkat which was a well organized community of
seafarers, but the archaeological situation does not allow the recognition of a smithy with
all necessary installations as a specific complex.
Moving southwarths and on to a slightly later period in time we meat the royal palace
(Königspfalz) of Tilleda in Thuringia (about AD 900 - 1190). This was a hillfort with
acropolis and bailey, a temporary residence of early medieval German rulers. Several
crafts were performed within the ramparts, some of them as duties fulfilled by serfs (e.g.
weaving in special workshop halls). Naturally, the blacksmith’s craft could not be absent.
Iron slags and tuyeres were found dispersed over all the area of the suburbium. Certain
accumulations were noted in neighbouring houses 1, 2, 15; 10, 18, and 226, 257 in the
western part and at fireplaces F 6, F 10, F 12, F 40. These may be dated to the 10th -
11th centuries but certain slag and iron objects were found in the upper layers of house 81
from the later finds horizon (the late 12th century). In spite of the existence of some slag
accumulations no complex workshop could be identified during the excavations (Grimm
1990, 95, Pl. 24 - 25, and 55a; slags were analysed by Waniczek 1987).


During the Migration period up to the 6th century AD many Germanic folks moved from
their sites and the territories were ready to accept new colonists (not raiders) - the Slavs.
Over 2 - 3 centuries Slavic tribes settled regions westwards up to the Saale river and
reached surroundings of Lübeck and Hamburg in the north-west.
The earliest evidence of Slavic ironmaking and ironworking remains indirect for the
6th nad 7th centuries AD with no finds or bloomery furnaces or smithy workshops, not
regarding occasional smithing slag in settlements. The 8th and 9th centuries saw a rapid
development of culture (incl. the coming of christianity from Byzantium) and economy,
followed by processes in social stratification and organization. Western Slavic regions saw
the establishement of a prospering pre-feudal state of Great Moravia (‘Megalé Moravia’
after Porphyrogenetos) met with political and military conflicts (successful wars with
162 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Fraconia, the catastrophy following the invasion of Magyars shortly after AD 900). The
metallurgy of iron in Great Moravia reached top levels (Pleiner 2000c, 276 - 277) and
the application of sophisticated technologies in the field of ironworking brought fruitful
results as in the light of metallographic investigation (e.g. Pleiner 1967d). Amounts of
iron produced enabled to introduce a value equivalent in form of pre-montary currency
- the axe-shaped bars in four standardized sizes (Pleiner 196l, ; Zaitz 1988; Bialeková
1981; Bialeková et al. 1999). Large-scale excavations of Great Moravian centres brought
examples of smithies as scenes of the blacksmith’s work.
One of them is Staré Město-Uherské Hradiště in central Moravia (a very important
princely and ecclesiatic centre of the 9th century AD). At Staré Město, ‘U Vı́ta’ site,
between churches 1 and 2, two sunken-floored smithies were investigated in 1976 - 1980
(Galuška 1992). Smithy feature 1 had, of an irregular ground plan (length 4.7m, 80cm
deep), and was equipped with a kind of an embanked domed furnace or kiln in the southern
wall the function of which was remained unclear, and with a horseshoe-shaped hearth in
the eastern bay (diameter 85cm). Eleven examples of PCB slag cakes were found but what
also is very important, a hoard of 23 iron tools was deposited in close vicinity. These too
concisted of bladed bark scrapers or small socketed hoes (o.c., 124 - 125, Figs 3 and 6 to
11). In this context smithy feature 5 (situated close by) is important. It had an irregular
sunken ground plan again with bays (max. 3m x 5m). The free-standing superstructure
must have measured some 4m x 5m. In the northern bay of the sunken floor a burnt-
red hearth lined with mixture of clay and sandstone gravel was installed (40cm x 60cm).
Another hoard of iron objects was buried in front of it: 32 broken axe-shaped bars of the
smallest size (lengths around 7cm). Another hearth was embanked in the eastern wall
(o.c., 126 - 127, 145, figs. 5, and 12 to 14, here Fig. 60: 2). It is worth of remarking that
the area of Staré Město yielded 190 pieces of axe-shaped iron bar of various sizes.
An entirely different smithy structure is represented by a slightly later workshop (de-
cline phase, around AD 900) in the acropolis of another Great Moravian leading centre
at Mikulčice (S Moravia), a fortified complex with 12 churches in marshes of the former
Morava river bed. The site also is known by the rich iron iventory, including a hoard
of axe-shaped bars and other irons buried under a rotary quern in church VIII. The
workshop in question (possibly a log building) marked by an rectangular clay-and-sand
dumped floor (m x 4. 2m) was two-roomed with what could be held for fore-room in the
south-east (2m x 4.2m, entrance not identified, Fig. 60: 1). The hearth was situated in
the SE corner of the main room, marked by burnt sandstones and burnt-red clay lining
and ‘light hammer scale’, hiding a shallow depression in the centre (200cm x 200cm over-
all the feature). Iron slags, about 100kg in total, PCB’s of 15cm dia. and 1kg weight
included, were concentrated in the hearth area and in some spots outside as well. Among
finds we should list are: a tubular tuyere fragment, slate whetstones, sherds of what might
have been a chimney hood and numerous iron objects - one axe-shaped iron bar, a sickle,
a spear, an axe, knives, a hook-key etc. Charcoal dust and ash cover most of the inner
floor (Klı́ma 1985). A relatively spatious smithy produced and/or repaired various iron
artefacts on behalf the upper class, residing in the acropolis of the Mikulčice centre.


The best information on the cultural milieu of the eastern Slavs is offered by the find at
Sarkel, a Khazar fort and later Russian agglomeration Belaya Vezha in the Don river
THE SMITHIES 163

Figure 60: Smithies in medieval settlement centres and suburbs. 1 Mikulčice, Moravia,
9th century AD, smithy in the bailey; 2 Staré Město, Moravia, 9th century AD, smithy
with a hoard of miniature axe bars; 3 Sezimovo Ústı́, Bohemia, early 15th century AD
suburb, smithy at a building complex. 1 simplified after Klı́ma, 2 after Galuška, 3 after
Krajı́c.
164 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
basin. A sunken-floored, apparently lightly roofed smithy feature 2 is known since 1950’s
(Fig. 61: 3). Special attention has to be paid to preserved organic elements of installation:
the wooden and leathern remains of a double air-inlet bellows (skin bag?) at an oval
hearth (100cm x 50cm, 20cm deep), a wooden beam as an anvil base, a wooden water
trough in the NW corner. In the opposit corner a small heap of what was originally held
for iron blooms (Sorokin 1957); the matter is, more probably, of PCB slags thrown out
into an ash-filled depression. No analysis has been published. The problem is of dating,
and, consequently, of the ethnic association. While Sorokin ascribed the workshop to the
Russian settlement Belaya Vezha of the late 10th century AD, Pletneva (1996) proposed
an earlier chronology, basing on the rediscovery of some metallurgical evidence (e.g. burnt-
red pit 28 covered with soils from the later northern fortress moat). She dates the above
described smithy (feature 2) and another one (No 5, sunken-floored, 5m x 5m, with slag-
filled pits) to the 9th century and considers the smithies as workshops working in a kind
of serail during the building of the Khazar fortress Sarkel.
Fedorov (1953, 117 - 122, fig. 50) described a sunken-floored smithy (8m x 4.5m) found
at the hillfort Ekimaucy, Moldavia, in the territory of the Tiver tribe. Two hearths placed
side by side in niches accessable from a sunken standing floor were the main installation
the fill of which contained iron tongs, a horned anvil, crucibles and copper wire. The
site was destroyed in about AD 1050 by the Pechenegi nomads. According to the author
of the excavation a human skeleton might have been the master smith killed during this
event.
Another two indications of smithy work were reported from the Hrynchug hillfort in
the Podol region (northwards, the 12th - 13th centuries AD, a stone-lined forge, 60cm in
dia. in front of a pit full or iron slag and charcoal, see Pachkova and Gopak 1981) and
from the Gat’ hillfort in the Orel region (the 11th - 13th centuries A). A forge hearth
is reported to have been placed in one of the block houses (see Rybakov 1948, 133). No
details are available.


Central Asia is outwith the scope of a European history. Nonetheless it would be pitty
not to comment a smithy workshop excavated at Pendjikent, an urban site of the 6th - 7th
centuries AD in the ancient Sogdiana. Room 109 measured 3m x 3m in area. Its western
wall has been removed but near the south-eastern corner a hearth with an air-inlet was
reported. The fill consisted of ash and slag occuring in large amounts. Another fire-place
was situated at the southern wall. Apart from slags, 24 small iron objects came to light
in this room. The site yielded a nice collection of clay twin-tuyeres for goat skin bellows
(Raspopova 1980, 36 - 37, fig. 23 and 33).

Smithies in medieval towns


In every town of the European Middle Ages, many blacksmiths, some of them highly
specialized (cutlers, armourers, locksmiths etc.), were settled. Reminders of their work
could be preserved in local names (‘Schmiedegasse’ - Blacksmiths’ Lane; Greatsmith Gate
at Nottingham, Kuznechnye vorota - Smith’s Gate at Kyiv). Towns like Toledo, Milan,
Nuremberg became famous for armour and weapons making, although no material evi-
dence has been left in their areas. Concrete finds of urban smithies are scarce. The reason
THE SMITHIES 165
is that the modern built-cover destroyed medieval levels or made them unaccessable. Af-
ter destructions of World War II a specific branch of systematical or rescue excavations
has been introduced - city centre digs (in German ‘Stadtkernforschung) during which and
during preparations for recent building activity smithing evidence could be traced or even
smithy workshops uncovered.
The 14th century Prague, the residence of emperor Charles IV, was a flourishing
metropolis. Charles extended it by the foundation of the New Town (Noua Civitas,
8th March 1348). A year before he appealed the craftsmen and among them the smiths
(except farriers and armourers) to settle in the new quarter and ordered that the work-
shops should be transferred to the rear of individual building plots not to disturbe by
noice. The centre of the new town was the Horse Market (now Wenceslaus Square). Doc-
uments say that in its eastern front 11 smithies were placed (among them 1 pin maker,
2 cutlers, 1 grinder) and in the western front 12 workshops have existed (among them
2 cutlers). Blacksmiths are called fabri but specialized branches are documented: cut-
lers (cultellifices, cultellatores, cultellifabri), nail makers (in Old Czech hrzebicnik), bridle
smiths (frenifices). Some of these craftsmen were wealthy and owned houses, sometimes
even imposing (domus magna) of Ješek alias Turšmid (Jeclinus Tursmid, Dursmido dictus,
1386).
Archaeological evidence concerning this ramified craft could be taken into account,
although in limited measure, during the building of the underground railvay station in
the cross-street in the middle of the Wenceslaus square, the former Horse Market (1967
- 1975). There was no hope to look for smithies (which, as said, were hidden in the rear
tracts of individual plots) but numerous PCB calottes emerged in the uncovered layers
as well as numerous iron artefacts, e.g. punch-marked knives and iron holders in the
wooden water-pipe system. Metallographic investigations of a set of these iron artefacts
(mostly knives) showed the top quality of the cutlery ware which aso was exported, e.g.
to Nuremberg (Huml and Pleiner 1991; Pleiner 1991a; Pleiner and Huml 1993).
Quite analogical data as for Prague are avilable for London. In the light of written
sources the specialized branches appear parallel with the smiths (14th century: cutlers.
armourers, cross-bow makers) which were settled mostly at the Ludgate and along the
axis leading to the St. Paul cathedral. The use of punchmarks was obligatory and to
develop noise by these craft at night was prohibited (Keene 1996). Iron slag and hammer
scale was ‘almost ubiquitous when occasional digs in the town were performed’ (Egan
1996, 91).
As to Britain, ironworkers were busy in medieval Nottingham (MacCormick 1996),
where armourers, arrowsmiths, cutlers (bladesmiths), farriers, grinders, locksmiths are
noted in written documents of the 14th - 15th centuries. The Greatsmith Gate in the
north of the medieval town is eloquent. Archaeological sources offer, however, occasional
finds of slag waste: Hight Pavement, Barker Gate (a PCB hearth bottom from a pit of
the 12th century).
Discussing the exploration conditions in medieval towns we feel to direct the attention
to Novgorod Velikiy in Northern Russia. Intensive excvations took place in the area of the
Kosmodem’yanskya street (Cosmas and Damian) which revealed medieval timber houses
and wooden street pavements and enormous amounts of iron artefacts of any kind which
were metallographically investigated by Kolčin (1959). Some evidence of ironworking was
involved (cementation crucibles, slag o.c., 12 - 16) but no smithy in the excavated area.
Indirect implications of that kind could be questioned in several localities from which
166 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
ironworking is not documented by written sources but some material evidence was cited
in the literature.
In the European south-east, in the ancient Bulgarian realm, passed the assimilation of
the Slav population with Turkish-Tartarian invaders and the christianity of the Byzantine
type was adopted. Outstanding urban centres like Preslav and Pliska flourished in the
9th - 11th centuries. In both, traces of blacksmiths’ work were recognized. At Preslaw-
Selichtche blackmith’s tools like tongs and wastes like slag were found together with other
iron artefacts (brandnew locks, cow-bells, mountings, crucibles, see Vazharova 1966) but
no workshop ground plan is known. From Pliska two smithies were announced (Bala-
banov 1981, 34 - 35, fig. 1; Doncheva-Petkova 1986) near the northern gate. They existed
presumably before the finishing of the fortification. The matter is of sunken-floored fea-
tures the description of which does not provide any easy survey and the identification of
individual workshops is practically impossible. Iron implements and fragments of copper
sheet should be accompanied with these finds.
In the northern part of Europe in the town of Ribe, Denmark, two cummulation of
iron slag from a smithy and a clay hearth shield (16. 5cm x 15cm) are noted by Bencard
(et al. 1979, 121 - 122) but no distinct feature was presented.


In the time period of the 13th - 14th centuries AD signs of smithies were reported from two
Italian sites. Rocca di San Silvestro near Campiglia Maritima, Tuscany, is an admirable
fortified urban complex, occupying a rock peak. It was investigated by R. Francovich. A
smithy was situated at the gate (Francovich 1993, 429 - 441, Figs 7 and 8). At Leopoli-
Cencelle, Viterbo in central Italy, a smithy worked near the main town gate (section III,
room III/H; no ground plan and dimensions are presented). Three furnaces/hearths filled
with slag and magnetic hammer scale are mentioned to have been sunk in the floor. A
depression (diameter 120cm and 20cm deep) comprised a smaller pit (30cm x 50cm, 15cm
deep) full of globular slag and hammer scale which indicate presumably the anvil position.
Ranuccio di Giovanni, the blacksmith, appears in a written document from AD 1220 (La
Salvia et al. 2001, 156 - 157). The evaluation of these finds is planned.
A special passage has to be devoted to the situation at Bocholt, Westphalia, W Ger-
many. The situation near the church reflects, in different strata, the evolution of a rural
site (villa Buocholt, horizon of the 11th and 12th centuries) to a town (the 13th and 14th
centuries). In all of the chronological phases ground plans of smithies were uncovered
(Reichmann 1984). During the village phase a sunken-floored smithy (house 44) with
four corner posts (3m x 6m) belonged to phase 3b (11th century). Iron slags were found
inside and outside the feature. Depressions and small pits might be traces of anvil and
bellows positions but non of them could have been interpreted with certainty, nor any
hearth. A file, a chisel, iron bands and door fittings were unearthed (o.c., 81 - 84, fig.
9). Another sunken-floored space inside the post-built house 45, 2.5m x 6m, was dated
to the 12th century (layer 3c). Remains of a domed furnace (3m x 3.5m) were found
outside the sunken place. Small slag pieces were present in the layer (o.c. 84 - 87, fig.
10). About AD 1200 (layer 4) this workshop was replaced by a post-built shelter (8m x
9m). This was probably an annexe building with a slag containing layer (o.c. 88 - 89,
fig. 11). Layer 5 (the 13th - 14th centuries) represented the urban phase of the site. The
smithy was a post-built complex consisting of two rooms (6.8m x 7.3m and 6.8m x 5.7m
THE SMITHIES 167
respectively) and a porch in the north where a rectangular pit called cellar was situated.
In the centre of the western room a large circular fired spot (diameter 2m) marked the
hearth which was opened to the west and supplied with air possibly from the east. An
anvil bed (diameter 50cm) was in a 1m distance southwards from the hearth rim. Slag
particles are reported to be present in the sandy soil (o.c., 90 - 93, fig. 12).

Let us turn our look to the very centre of Europe, at time the principality and later the
kingdom of Bohemia. Intensive works during the rescue ‘Stadtkernforschung’ revealed at
Žatec, NW Bohemia, several smithies which are not yet published with all documenation
but introduced by Čech (1998) in a prelimnary report. From levels uncovered during the
rescue works which date from times before the town achieved a royal statute came to light
seven smithies: No I was sheltered by double walls and comprised two burnt-red hearths
and pottery of the 10th century AD; smithies II to IV were sunken-floored workshops
equipped with semi-circular clay-lined and burnt-red hearths. Number V is said to have
contained 5 hearths in an L-shaped arrangement (the late 12th century AD). All of these
workshops contained PCB slag in their fillings. In smithy VI, of the same date, were 3
superimposed hearths and elongated and cauldron-shaped reheating installations. Finally,
the Chelčický Square excavation brought a feature (VII) which was labelled as a cutler’s
workshops with finds of several knives. No doubt that the final report and publication
will throw more light on the appearance of early medieval blacksmiths’ workshops.
One of the best illuminated and documented smithies was that having been uncovered
at Sezimovo Ústı́, S Bohemia (Fig. 60: 3). This was in a part of the town suburb (called
Nové Město) situated on the left bank of the Lužnice river (systematical excavations have
been carried out by M. Richter and R. Krajı́c since 1960’s). It should be noted that exact
dating is available (AD 1419 terminus ante quem). In that year the population abandoned
the site and moved to nearby Tábor at the beginning the Hussite movement. The com-
plex in question (estate V, excavated by Richter in 1960-1965), a trapezoidal ground plan
(13.2m x 10m x 9.6m x13.8m, 37. 8 m2 ), consisted of five stone-walled rooms; those to
the east and south were dwellings and stores, the NW room was the smithy and was open
to the east (and sheltered). In the SW corner was a squarish hearth (140cm x 140cm)
with stone-lined rims, open to the west, into the workshop room. Access to this hearth
was limited two brick-made plinths. The hearth bottom was doubed of small stones and
clay and was fired red. A depression sunk below the rammed bottom (20cm x 20cm) in
the hearth centre was apparently the lowest part of the fire-pit proper and filled with ash.
There was a pit (diameter 40cm, 10cm deep) in front of the hearth, in a squarish area (3m
x 3m) demarcated by grooves, which could have been the bed for a wooden anvil block.
600kg of iron slag and both globular and lamellar hammer scale covered the floor of the
workshop and extended outside. Scrap iron, horseshoes, 1500 nails, mandrels, bars, band
iron, locks illustrate the work of a universal blacksmith and farrier who had mastered sol-
dering with copper as well. A purse containing 19 Prague groschen was buried under the
floor (Krajı́c 1993; 2003). The workshop was owned by a relatively well-off master smith
who supplied the customers with necessary ironmongery and rendered repair services.


168 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
This excursion over the smithies in medieval towns leads us to the North of Europe, to
mid-Norway. Trondheim, the early medieval Nidaros, was not only the political capital
and ecclesiastic centre but also an important trading post. Many crafts satisfied the needs
of the inhabitants. An insight into the activity of metalworkers has been provided by ex-
tensive rescue digs at the sites of Mellager and Library which were rendered necessary
by the construction of commercional and public buildings. The Mellager site occupied
the estuary of the Nid river into the Tronheim fjord, the Library site is situated a little
southwards on the left bank of the Nid river, at the medieval harbour. Metalworkers
were active there, both founders and blacksmiths who operated in sheltered spaces. The
production of the early 11th century left some crucibles, clay mould fragments and ‘tiny
quantities of smithing slag’. Possibly, the same craftsmen mastered non-ferrous and fer-
rous technologies but on small scale. Itinerant artisans might have been active during this
phase as well (McLees 1996, 123 - 125, Figs 4 to 6). In course of the 12th - 14th centuries
a metalworkers’ quarter arose to the north on the estuary (Mellager site). According
to McLees (o.c. , 128) the workshops were simple log cabins with sunken hearths lined
with stones or clay (diameter 50cm to 70cm, 30cm deep, ibidem fig. 10). About 600kg
of slag represented the waste. Some PCB calottes were chemically analysed by Espelund
(1992) who is inclined to see in these bottoms traces of refining blooms imported from
the Budalen bloomeries in these finds. Jakobsen (quoted by McLees, 178), on the basis
of slag analyses from the medieval Tonsberg, would prefer an interpretation of PCB’s as
waste from welding operations in which phorsporus-rich fluxes were used. An inflow of
metalworking craftsmen followed during the subsequent centuries. However, the mid-14th
century saw the decline of the role of this part of the medieval town, possibly influenced
by the plague epidemic of the 14th century. A rennaissance can be traced in the 15th
century. The archbishop’s residence, close to the cathedral, housed armourers’ and mint
plants (McLees, o. c., 130 - 132, fig. 15). The timbered weapon workshop, dating from
c. AD 1500, consists of two main rooms (3m x 5m, and 4m x 5m, subdivided by parti-
tion walls). Three hearths functionned in that structure, a large one (‘A’, a forge, in the
eastern room corner, stone built, aparently elevated), and smaller fire-places (‘B’ and ‘C’
in the other room and compartment). These were sunk in the tiled floor and were used
during other heating operations. Cross-bow bolts and lead projectiles are supposed to
have been manufactured in this late workshop which is so different in comparison with
earlier smithies.

Rescue excavations in the centre of the old town of Zug (ancient Tugium) in Switzerland
brought to light at least seven places with finds or even accumulations of smithing slag
(a great deal of which were plano-convex calottes) and, in addition to that, relics of two
workshops dating from the late medieval to early post-medieval periods. The one traced
in the Untergasse 18/20 in Sust was badly disturbed by the rebuildings of the later house
and left several remains of hearths from different periods (the 12th/13th to 16th/17th
centuries) and vaguely described smithing slags (Senn-Luder 1998, 129 - 132, fig. 15).
The other smithy, Ägeristrasse 8, was the ruin of a post-built structure (c. 5m x 6m)
with a rectangular stone-built and originally elevated hearth in the NW corner (100cm x
50cm, destroyed). There were pits with hammer scale, one of them (G 3) combined with
a protrusion that was apparently held to be the anvil position. The date is the 15th -
16th centuries (ibidem, 132 - 137, fig. 17).
THE SMITHIES 169
Medieval rural smithies

We will attempt to show that medieval towns supplied the country with specific items
of ironmongery; simultaneously, facts will be presented that smithies in villages not only
maintained and repaired iron objects necessary for rural everyday life and work but some
of them also produced various kinds of artefacts on their own, including cutlery.
The lines below will present European rural smithies as published in the literature.
First, sunken-floored features i.e. placed in houses or shelters with floors below the walking
surface will be treated (in their approximate chronological than territorial order).
Kosel near Rendsburg-Eckenförde in the German part of Jutland yielded some sunken-
floored late 8th to 9th centuries AD houses which are a little atypical for the region.
No 62 was a smithy, rectangular in plan (4.2m x 3.3m, 60cm deep, Fig. 61: 1). In the
SE corner a heap of stones and fired clay (100cm x 100cm) marked a furnace denoted
as domed installation; in the SW corner another stone accumulation was the elongated
second hearth (200cm x 100cm). In the centre of the sunken plan there was a pit filled
with charcoal on the edge of which a large stone might have served as an anvil. It was
surrounded with iron slag (24 pieces comparable with the PCB’s). Forty iron fragments
of that feature were taken from the fill (Meier and Reichstein 1984, 118, pl. 10: 1).
The smithy of Lebedka was a little later (the 8th - 10th centuries AD) but in an entirely
different cultural milieu (ancient Slavic Russia). Feature 2 is a remarkable plant sunken
in the sandy soil the ground plan of which forms a number-of-eight shape covering the
area about 5m x 4m in total. The western part (3m x 2m, 80cm deep) involves another
sunken square-shaped pit (additional 80cm of depth) - the smith’s stand point paved with
wood (1.8m x 1.8m, Fig. 61: 2). The master could operate an elevated hearth made in
front of him of burnt-red compact clay (stone-lined, dia. 70cm). The rear part of the
hearth formed a small wall (50cm high), apparently as a protection of the bellows. Neither
tuyere nor air-inlet is mentioned in the published report but the bellows must have been
situated in the fan-shaped space behind the hearth (1.5m x 2m, more than 1m deep). The
fill consisted of ash, charcoal and iron slag which also was thrown out and has been found
outside the hut as well (Nikolskaya 1954, 100, fig. 42: 1; eadem 1957). It is not known
what kind of superstructure covered the sunken-floored space but a timber building comes
into question. The deep stand of the smith is worth of stressing since it also was applied
in other cultural provinces up to modern times as ethonolgical examples show.
Turning back to western Europe a brief remark should be made about a smithy from
the 9th, beginning the 10th century AD at Liestol-Munzach/Röserntal in Switzerland
which was presented by Tauber (1992, 28 - 29, fig. 3). The sunken plan was relatively
shallow (4m x 8m) with four posts in corners. In the middle of the western wall was a
ruin which was denoted as a large fire place. The bellows should have been operated from
outside the hut. About 2 tonnes of iron slag (incl. PBS’) were found in and outside this
feature.
In western Hungary irregular pit-workshops with hearths (dia. 30cm) and iron slags
are mentioned at Csatár and Csongahegyhát in Zala region and at Várpalota by Valter
(1981) and Gömöri (2000, 332) but no details are presented. The dating: the 10th - 11th
centuries AD.
The 12th century is the proposed date supported by some coins for the sunken-floored
smithy (about 50cm deep as excavated) found at Hidészég. Sopron region, western Hun-
gary. It measured 3.5m x 4m and in shorter sides were posts for supporting the ridge
170 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 61: Medieval rural smithies. 1 Kosel, N Germany, Viking period; 2 Lebedka,
Russia; 3 Sarkel-Belaya Vezha, Russia, 9th/10th century AD; 4 Stöng, Iceland, Viking
period; 5 Haus Rhade, W Germany, 14th/15th century AD; 6 Mutějovice, Bohemia, 13th
century AD (two subsequent features). 1 after Meier, 2 after Nikolskaya, 3 after Sorokin,
4 after Capelle, 5 after Sönnecken, 6 after Pleiner.
THE SMITHIES 171
beam. It was transformed to a smithy secondarily and the hearth (diameter 30cm) was
set into the remains of destroyed domestic oven or kiln. Smithing slag (about 50 examples
of PCB cakes) and numerous horseshoe-nails were found so that the interpretation as a
farrier’s workshop appears as plausible (Gömöri 1995, 187 - 201, Figs 6 to 9; a post hole
near the entrance is presented as a binding stake for horses in the reconstruction).


Two sunken-floored smithies of the late 13th century were excavated in 1963 and 1965
at Mutějovice near Rakovnı́k, western central Bohemia (Pleiner 1969b). An enormous
accumulation of PCB slag appeared in an area c. 10m x 10m, not regarding numer-
ous stray finds in the vicinity; 14 settlement features from the 11th/12th to 14th/15th
centuries were uncovered there of which two smithy workshops are archaeologically sig-
nificant. Both are dated to the the 13th century but stratigraphically they appeared in
superposition, smithy A being the earlier than smithy B damaged by a corner its rim. In
fact, the situation was more complicated since smithy A was slightly cut by other two
pits, Nos 6 (a shallow depression, damaged by smithy B) and 5 (presumably a domestic
stone oven, possibly contemporary with smithy B).
Smithy A (feature 1) was a rectangular pit with the northern side slanted (2.6m plus
6m slanting x 4m, 40cm to 50cm deep, with depressions reaching up to 106cm; sloping
walls). It is assumed that about 40cm to 50cm were gradually ploughed out and should
be added to depths. In the eastern bay was another depression reaching the depth of
126cm from the modern surface. This was apparently the stand of the smith who faced
to the slanting wall at which a clay-lined and burnt-red hearth was placed (external
diameter 50cm, internal diameter 30cm, preserved depth 5cm - 6cm); the fired lining was
permeated with hammer scale. The bellows must have been situated at his left hand (no
direct indication), the anvil at his right hand. In the sloping wall was a small sunken
pit (60cm a 30cm, 10cm deep) which could be held for a sign of an ancient anvil bed.
The fill of the feature yielded 723 pieces of PCB calottes (cca 11cm in dia.) weighing
483kg in total, apart from ash and charcoal. Further finds included: fragments of tubular
clay tuyeres, irons which escaped the evacuation, unfinished bone handles for knives,
and pottery sherds from 13th century high storage vessels. The iron objects comprised
a fragment of the upper part of a plougshare, a cylindrical iron padlock (undoubtedly
imported from a town but repaired here and equipped with a new iron socket), knive
fragments, nails, an arrowhead, and wire. The local smith manufactured some kinds of
iron objects and repaired those valuable enough in his universal practice.
Smithy B (feature 7) cut slightly the rim of smithy A and was situated northwards.
The ground plan was a little trapezoidal with rounded corners (5.9m x 4.3m). The present
depth was 70cm - 80cm but in the northern part a depression (1.2m x 1.2m at the bottom)
reached the depth of 110cm. Again, it could have been the stand for the smith, although
the hearth on his right hand was relatively distant (1m). The hearth was a depression in
the sloping pit wall (126cm x 94cm, with fragments of fired and slagged clay lining). The
bellows must have been placed in front of the smith on the small flat platform. The anvil
position is not known. The filling contained charcoal and 430 pieces of PCB slag cakes
(211kg in total) and nails, animal bones and pottery sherds of tall storage vessels. The
date: the final 13th century. The successor of smith A declined to renew his workshop on
the old spot and choosed a more solid terrain slightly in the north. It is possible that he
172 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
worked there for a shorter period. None of the two smithies were destroyed by a catastro-
phe, they were peacefully abandoned; valuable things made of iron were evacuated. No
indication exists as to the nature of the superstructures over the sunken workshop spaces;
they could have been timber or log houses. Fig. 61: 6.

Freestanding structures, in our case the smithy workshops, can be recognized under
favourable soil conditions at the site, either in preserved or slightly eroded historical
surfaces or floors.
An early example (the 6th - 7th centuries AD) is represented by farm 6 in the hamlet
of Develier-Courtételle, Jura Switzerland. A hearth was uncovered (90cm x 80cm, 10cm
deep) in the western part of a post-built structure. 4.13 tonnes of slag were collected (99
500 pieces and fragments), among them 1370 PCB calottes which were mineralogically
analysed (Eschenlohr 1998). The authors suppose that the smiths both refined imported
blooms and forged objects. Was the quite extensive ironworking activity in the Courtételle
‘Weiler’ undertaken on behalf of its inhabitants or was it a matter of the production of
iron objects of every kind for broader distribution? The evidence that we have at our
disposal at the moment does not enable us to answer this question.
A freestanding smithy bordered by stakes and occupying an area 11m x 10m was partly
excvated at Nemeskér, Retrejáró alja (Gömöri 2000, 112 - 12., Figs 62, 67, an 71). Apart
of smithing slag cakes iron knives and other artefacts were found, among the a peculiar
kind of small tongs, and tubular tuyeres (o.c., fig. 63, and 74 to 76).
In central Germany, at Assum near Hildesheim, an 11th century estate was found with
rectangular free-standing buildings: the main dwelling house, a separated kitchen and a
light post-bult shelter, one side open and facing a well - a smithy (6m x 4m, see Barner
1935, fig. 2: 1). A hearth of c. 100cm in diameter was surroundd by an ash layer from
which iron artefacts, especially nails and horseshoes and whetstones were unearthed. No
slag waste was mentioned but a ‘slag-like sintered bloom’ `Roheisen, sog. Luppe, o.c.
122). The find cannot be controlled; it could represent a bloom or PCB calotte. The
smithy of Assum undertook the manufacturing and repairing of domestic equipment and
provided a farrier service.

In spite of the rather chronological nature of this survey of smithies the rural settlement
complexes of medieval Scandinavia must be treated as a unit. These involved smithies
as well. We have already had the opportunity to discuss the site of Vorbasse (Jutland,
Denmark) in connection with Romano-Barbarian workshops. The site was inhabited
in the Viking Middle Ages as well and smithies were among the installations that the
population required (Hvass 1978; Müller-Wille 1983, 236 - 238, Figs 6: 5; 8: 2). Four
smithies were referred to in the eastern complex of estates with longhouses (no details
available) dating from the phase of the 8th - 10th centuries AD. The western group of
farms, dating to the later phase of the 10th - 12th centuries, involved four smithies, two of
them as separate house features (finds of iron objects and iron bars), and two as built-in
workshops inside the longhouses, one in the western end of such a structure, the other
THE SMITHIES 173
in the middle of the building. The working of both iron and bronze were reported (see
Müller-Wille, o.c., 236.
North-westwards from Ribe (Jutland, Denmark) a similar Viking Age settlement was
investigated at Saeding. Inside and outside the longhouse III which was the residental
building, a smithing hearth and hearth shield and anvil stone were found, as well as iron
slag, scrap iron and a bundle of iron rods or bars for rivet and nail making. House II was
a workshop equipped with a hearth, protection shield of stone and stone anvil. Iron scrap
was found. A smithy was located in an out-house in the north of the excavated area. It
is not clear whether the two latter smithies worked at the very same time (the settlement
was multi-phased) but two smiths could be useful within a relatively large habitation
complex.
A different model of farmsteads can be met on Shetland isles, Iceland, and in Green-
land. These were self-sustaining, more or less isolated units consisting of several buildings
incorporating separate smithy features. At Jarlshof, Shetland, such a rectangular smithy
(I B) was situated between two longhouses (I and I C, see Müller-Wille 1983, 239, fig. 10:
1, after Hamilton).
The situation was similar at Stöng, Iceland (ibidem, 239, fig. 10: 2). The stone-walled
workshop (8.5m x 2.6m) housed a hearth, a stone for fixing the anvil, a stone-hewn water
tank (Capelle 1980, 428 - 431, here Fig. 61: 4). There are references to similar situations
at Hraftnseyri, Hrauntunga and Whórarinstadir. The Stöng smithy mus have existed
before the eruption of the Hekla volcano in AD 1084.
Before dealing with further two cases of estate smithies it is necessary to refer to
workshops at Reykjavı́k, the modern capital of Iceland. The first, older, dates from the
second half of the 10th century AD and comprised a stone-lined hearth with a stone
bellows protecting shield, housed in a building with stone-sustaining walls. 11.5kg of slag,
charcoal and burnt bones formed the waste material (Nordahl 1988, 112). The later smithy
(12th century) yielded 2kg of slag, hammer-scale and fragments of iron objects (ibidem.
No details are presented. The ‘raudasmide’ from Ormsstadir (a stone-walled house of
8m x 2.1m) was a medieval bloomery (ore store, slag, furnace and anvil, according to
Fridriksson and Hermans-Audardótir (1992).
The model of a farm served by its own smithy appears on the western coast of Green-
land, namely at Sandnes - ‘The Farm beneath the Sand’ (Müller-Wille 1983, 239, fig. 11:
1, after Krogh; Buchwald 2001, 51, figs 44 and 49). There were two byres and the main
residence building and, inbetween, a small stone-walled house or shed 7 (2m x 2m) called
a smithy. Buchwald (o.c.), after having analysed the PCB slags was sure that these were
‘typical purification slags’ from refining imported Norse blooms. The Sandnes farm dates
from the period AD 1200 - 1250.
Across the ocean to the west the Vikings settled at a bay in New Foundland, America,
At L’Anse aux Meadows a group of turf-walled buildings represents the dwelling area,
while accross a river, in the west, traces of four ship sheds were visible in the bay. A a
charcoal pile and a workshop denoted as a smithy also came to light (Müller-Wille 1983,
239, fig. 11: 2, after Ingstad). The workshop rendered services in terms of keeping iron
implements and tools used in the agriculture, fishing and hunting and in maintaining the
iron mountings and rivets of the ships.


174 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
To conclude the topic of rural smithies it remains to deal with those directly or topograph-
ically connected with bloomery ironworks or with working the products of iron smelting.
From Kamionka Nadbuźna, Poland, Rauhut (1956, Figs 2 to 8) published a groundplan
of a large rectangular house or shelter (4m x 8m) the limits of which were marked with
discarded PCB slags. There seems to have been hearth and a pit filled with slag inside.
Features denoted as reheating hearths were situated in the northern vicinity of the house
(12th century AD).
The site of B 85 at Eschenburg-Wissenbach in an iron producing district of the Lahn-
Dill river basins in western Germany deserves attention because a smithy involved in the
iron processing was uncovered there. The dry-stone-walled building (6.3m x 3.6m) was
probably open against the west where a porch might have been set up. Two stone hearths
with sunken fire pits were identified, one in the SW and another one in the NE corner.
Inside the bulding 640 kg of iron slag were found, up to 4 tonnes in total, including the
surrounding area. Some iron objects like knives and nails came to light as well. No
hammer scale was among the waste. Willms (1995, 76 - 77, fig. 14) classifies the feature
as a smithy refining iron blooms from ironworks in the vicinity. The chronological times
span is given as 12th - 14th centuries.
Haus Rhade near Kierspe and Lüdenscheid, Märkischer Kreis in western Germany is
the locality where one of the earliest charcoal blast furnaces (13th century) was excavated
in 1967. Three years later, in 1970, another feature was uncovered at a distance of about
20m from the furnace but on the opposite bank of the Kierspe brook. This was a rectan-
gular structure of drybuilt stones (6m x 4m, internal dimensions 2.8m x 4.6m), with an
entrance in the eastern corner (Fig. 61: 5). Hearth F 1 (60cm x 70cm) filled with ‘bloom
fragments and reheating slags’ was located in the centre of the room. According to the
pottery finds the installation dates from the 13th century. The second fireplace F 2 in the
northern part of the room (65cm dia., 25cm deep) containing no traces of smithing activ-
ity was later in date (Siegburg pottery of the 14th - 15th centuries). At the northern wall
there was what is called debris pit (50cm x 75cm, 40cm deep) which was filled up in the
15th century. The long existence of the building is indicated by numerous pottery finds
(from later ‘Kugeltopf’ to Siegburg ware); iron objects include mountings, 1mm thick iron
sheets, hooks, nails, rings, horseshoes, knife blades; a dagger and a sword. The smithing
hearth proper was placed at the northern wall outside the room. A stone plate (60cm
x 65cm, 12cm thick) bordered by by vertically set stones so that the dimensions were
75cm x 100cm covered the base. The installation was burnt-red surrounded with a layer
containing iron objects and pottery of the 14th - 15th centuries. No slag is mentioned in
the description. At the opposite southern part another oven adhered from outside to the
wall. This was semi-circular and lined with stones (300cm x 150cm); possibly a baking
oven. Numerous slag heaps in the vicinity of the ‘Flossofen’ and the smithy were levelled
during the pasture cultivations. Different kind of ironmongery from fittings, domestic
gear over horseshoes up to weapons show that the smithy rendered services for the Haus
Rhade estate during the 12th/13th to 14th/15th centuries. In fact, its activity was not
connected with the blast furnace (Sönnecken 1977, 27 - 38, figs 7 to 16).

Medieval castle smithies


Feudal castles as residences of nobles controlling and governing vast lands were above
all fortresses playing role in the defence of territories and in guarding important com-
THE SMITHIES 175
munications but they also served as administrative centres. Castles were supplied with
subtenance and necessary goods from outside (dependent villages and town markets) but
to certain extent they were self-contained units and as such they had to be able to main-
tain and keep the buldings and movable stock. Craftsmen as masons and stonecutters,
carpenters, joiners, slaters and blacksmiths were an integral part of the castle population.
The smiths repaired damaged and used iron artefacts, carried out the farrier work and
sometimes produced new items e.g. arrowheads, crossbow-bolts and the like. Smithies
have been discovered in several castles during restoration work, rescue digs and system-
atical excavations.
A rescue operation took place in the recent third courtyard of the castle at Jindřichův
Hradec, S Bohemia, in the space between the main fortification and outer defences. A
part of a castle smithy belonging to the 13th century was excavated. The sunken-floored
feature (4m x 4m) was equipped with a brick-set fired hearth (diameter 50cn) surrounded
with a quantity of slag (about 50 PCB’s). Preliminary information was provided by T.
Durdı́k (1990).
The excavation of the castle of Lelekovice near Brno, Moravia, started as a rescue
dig but it developed to a large systematical investigation. It should be underlined that
the castle of Lelekovice was entirely buried, not visible on the surface; the excavation
uncovered its roughly circular ground plan anew. The results were finally published in a
monograph by Unger (1999). A concentration of iron slag adhering to the fortification
rampart was uncovered in the eastern part of the castle and subsequently the ruins of
a stone-an-clay hearth (feature 301) were observed occupying a space 2.3m x 0.6/0.8m
(Fig. 62). Apparently this was a substructure for an elevated fire place. Pit 307 with a
levelled bottom (110cm x 120cm, 20cm deep) could have been the bed for a wooden stem
with fixed anvil. Post-holes may have indicated the bellows position south of the hearth.
Other post-holes and grooves show that the working place must have been covered with
a light shelter (Unger 1999, Figs. 63 and 63). A sunken vessel (312) was found close to
the hearth which was filled with lamellar magnetic hammer scale. The slag was of the
PCB type and was investigated by Mihok et al. (1997). No blacksmith’s tools were left
in the workshop area but smithy tongs with short jaws were found in the ruins of tower
H (ibidem, fig. 65). Iron objects (tools, fittings, weapons) were found in the layer of the
castle excavations. An iron chisel with a punch mark deserves attention (o.c., fig. 120:
1): this was without any doubt imported.
Racia̧że near Tuchola, SE Poland, was a govenor castle (the 12th - 13th centuries). A
smithy workshop was situated in the dwelling structure of this fortification (Kowalczyk
1976). Charcoal, slags and blacksmith’s tools are reported (a block anvil and tongs). A
rather destroyed site was found at Siedla̧tków near Poddȩbice, W Poland. Archaeological
excavations (1965 - 1966) of a hillock lead to discovery of a totally burnt wooden donjon
of a knightly stronghold of the 14th century (Kamińska 1968). A smithy workshop is said
to have been situated in the basement where ruins of what is held for a domed furnace
were elaborated. No slags and other kind of waste are mentioned but there were smithing
tools: tongs, a hammer, a nail punch and 1120 iron artefacts of all kinds from nails,
sheets and fittings over knives, sickles, horseshoes up to spurs, crossbow bolts, fragments
of plate armour and a helmet. An iron crucible for melting non-ferrous metals and bronze
sheet complemented the metal inventory. 52 iron object were submitted to metallographic
investigation (Nosek 1968).
Three chronological phases are indicated in the existence of an ironworking site at the
176 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 62: Lelekovice, Moravia. Sheltered smithy at a rampart (traces hatched) in a


castle of the 14th century AD. After Unger.

Manor of Alsted, Surrey, England (Tylecote 1986, 191, 192, fig. 131, after Keteringham).
A small slagged bowl hearth (diameter 25cm, depth 28cm) was located there around AD
1250 - 1270. Between AD 1270 - 1340 two tonnes of slag and other wastes were deposited
in an adjacent dump. In the third phase (AD 1395 - 1405) a stone substructure (about
50cm high) was built where a rectangular stone-lined hearth was installed which was pre-
sumably blown from one of the sides. In front of the hearth a stone supposedly served
as an anvil support; the stand point of the smith must have been there and not on the
opposite side of the hearth on the elevated platform, as proposed in the reconstruction.
The workshop must have been sheltered. No slags are mentioned in connection with the
last phase.


Smithies or traces of their work have been found in several medieval castles but they date
from final or even post-medieval chronological phases. Trosky castle which dominates a
THE SMITHIES 177
great part of north-eastern Bohemia regions was founded in the 14th century but held by
Swedish troops still during Thirty Years’ War. Possibly during the 16th - 17th centuries
A smithy was installed at the northern lower fortification between towers which crown
two rocky peaks perhaps during the 16th - 17th centuries. Rescue excavations took place
in 1997/98 (Prostřednı́k and Hošek et al. 1999. A hearth (40cm dia., 10cm - 15cm deep)
was sunk in the floor and backed by an L-shaped stone-and-clay wall in the east (160
cm) and south (80cm). The bellows must have been situated behind the eastern face.
However the air-inlet itself could not be properly identified. Three PCB slags (0.8 - 1kg)
were analysed by Hošek (o.c. 31 - 32): a great amount of it must be buried on the
northern forested slope of the castle hill. Iron objects found included nails, horseshoes,
hinges and arrowheads. The function of the smithy is undertstood as providing services
to the garrison of the fortress. No refining of imported iron blooms comes into question
at this site and at that time.
Smithies in three castles have been reported from eastern Slovakia. The stone structure
of a rectangular room (4.8m x 4m) has been uncovered set against the outer fortification
of Šariš castle close to the bastion 10. Two destroyed stone furnaces (80xm x 35cm,
containing three iron cannon-balls of 5 to 10cm diameter) were found in the eastern corner
of this feature and a fire place came to light in the opposite southern corner (diameter
60cm). Iron slag was dispersed around on the floor. There were nails, cross-bow bolts,
a spear point, horseshoes but also crucibles and tongs for casting lead bullets (diameter
2cm to 3cm) among the iron finds. Two extended inhumations excite the interest (at
the southern side of the room, Slivka 1978, 235-236, fig. 8). A stone-walled room with a
trapezoidal ground plan (3.3m x 5.4m x 3.8m x 3.8m) was found in the western end of of
the lower courtyard of castle Kapušany near Prešov (E Slovakia). It was a smithy with an
irregular quarry stone-and-brick hearth at its southern wall (150/190 cm x 173/185cm)
backed by a low wall and a central pillar. PCB slag and iron nails were spread on the
tiled floor. Dating: 17th century (Slivka 1978, 245 - 246, fig. 11). The smithy in the
fourth courtyard of an imposing castle of Lubovňa (NE Slovakia) which measured 4m x
6m dates from around the same period. PCB slag cakes and fragments of fired hearth
lining turned up among the waste and horseshoes, as did nails, a fire-steel and various
fragments of fittings. Date: the 16th - 17th centuries (Slivka 1978, 244 - 245).
The main role of all of those castle smithies was apparently that of maintenance and
occassional provision of required iron objects.

Monastic smithies
A similar but simultaneously slightly specific unit was the medieval monastery, richly
endowed by early medieval noble. In time these monasteries undertook ecnomic activity
on a grand scale also making use of the not unimportant means of taxes and tributes.
Moreover, some ecclesiatic orders pursued ore mining and metallurgy, e.g. the Cistercians
and Carthusians in France (Verna 1995) and Switzerland in particular (Eschenlohr 2001,
144 - 147). Many monks pursued artisanal crafts, among them metalworking and iron-
working as the example of records presented by Theophilus Presbyter (possibly Rogerus
of Helmershausen) show. He compiled a manual ‘Schedula diversarum artium’in the 11th
century AD.
Anyway, smithies in cloisters fulfilled, in general, similar task to those in the castles.
The Benedicine monastery of St. Gallen, N Switzerland, is represented by a reconstructed
178 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
ground plan as it was around AD 820 (Westermann’s Atlas zur Weltgeschichte 1963, map
53: III). Apart from the brewery and bakery artisans’ houses which were loccated in the
southern front of the complex, a swordsmith, a shieldmaker and a blacksmith worked in
relatively small rooms (no scale given).
Archaeological evidence concerning monastery smithies is unfortunately scarce. There
are in effect only three sites worthy of mention. They are atypical and even extraordinary.
Scott (1990, 161, Figs 6.5.5 and 6.5.6) describes, after Fanning, a part of a monastic
settlement at Reask (Co. Kerry, Ireland) dating from the 5th to the 10th centuries AD.
Small coupled circular stone cells C and D were connected with a doorway. Room D,
with an entrance in the north, revealed three pits which were interpreted as bowl fur-
naces and/or reheating hearths for bloom operation. Cell C was badly disturbed showing
remains of burnt-red pits, charcoal, slag, crucibles and perhaps tuyere fragments and
suggesting the role of a general workshop or forge. The central circular fire installation,
covered with stones on the bottom, took an area of 2m x 3m. A slag dump was deposited
outside the cells in the east.
Tylecote (1986, 190, fig. 129; 191) reproduces, after Higgins, a schematical ground
plan of a relatively large monastic smithy of the 12 century AD from Waltham Abbey
in Essex. It is a stone-walled internally divided room (about 8m x 15m) with two stone-
and-brick-built bases for hearths and several pits, store and ash deposits. Another small
pit served as a hearth for smelting lead. It seems that it was a multifunctional workshop
where, occasionally, iron ore used to be smelted. No slags are described.
A different configuration is offered by the Cistercian abbey at Fontenay, France, which is
now known as one of the centres of iron production (the 12th - 13th centuries). The monks
ran iron mines (at ‘Les Munières’, north-westwards from the monastery), bloomeries and
smithies and the product sold on markets. A 50m long stone building called forge is still
preserved south of the abbey. A water channel (supplied with water collected by dams)
passed by and served for hydraulical devices like a hammer-mill which is presumed to have
been located in the western room D (10.2m x 10.6m). The adjoining room C, of about
the same dimensions, was two-storeyed and is being interpreted as the forge where blooms
were reheated. As to the wastes, a slag heap called ‘crassier des moines’ was situated
south of the forge (Benoit, sine anno). The technological and economical extremity of
this place is evident.


The presence of ‘customary’ smithies in cloisters was presumably the rule. Unfortunately,
no excavated example can be presented. In 1958, K. Bielenin and the author undertook
an occasional survey on the slopes where the Tyniec monastery near Cracow rises. Typi-
cal PCB slags which had been thrown down from the cloister were found but the smithy
proper has not been located.

Smithies at mines
A specific kind of workshop is represented by medieval and post-medieval smithies in-
stalled at gold and silver mines where hard rock was combined with ore veins. The
miners chiselled it with iron-and-steel picks which became damaged within a relatively
short time and needed reconstruction and repair. Therefore, operative help must have
THE SMITHIES 179
been available in special smithy workshops in short distance or even at the adit entrances.
The masters in these smithies must have been skilled and experienced since the tools
they took care of were sophisticatedly manufactured iron-and-steel artefacts as numerous
metallographic studies show. Several mine smithies could be traced or excavated and
investigated.
The polymetallic resources (copper, lead, silver) that were mined in the southern
Schwarzwald (SW Germany) have produced traces of mine smithies at Sulzburg: five
at the site of Riestergang, and two at Himmelschergang (11th - 12th centuries according
to pottery finds, the 13th century according the radiocarbon dating). A smithy appeared
in the upper level of an abandoned and filled up mine at the first site; PCB slags were
found and a heap of that very waste was identified nearby (Goldenberg 1992, 240 - 24,
fig. 5; idem 1993; Zimmermann 1993, 220 - 222). At St. Ulrich-Bickenberg, in the same
region, several mine smithies were uncovered and mentioned by Zimmermann (1993, 216,
fig. 10) and Goldenberg (1992; 1993). In one case a burnt-red oval hearth (45cm x 30cm)
was mentioned which must have been blown by air from the south (tongs were found in
the access groove); PCB slag heaps were reported.
More complex information comes from the region south at Brandes in Oisans mountains
(near Hues, Isère, France) where silver mines were exploited during the 13th century. An
extended miners’ settlement was located on a platform at an altitude of 1800m. The
houses facing the south were a little sunken-floored, the walls erected of stones with
quartz-permeated clay (Bailly-Maı̂tre 1987). The same building technique was used in
the case of four smithies which yielded valuable data. Two of them (B 68 and B71b) were
discovered in the eastern part of the village of Brandes, close to one of the mining zones.
B 68 was a slightly trapezoidal room (2.5m and 3m wide and 4m long of internal space).
A rectangular hearth was leaned on the western wall (the stone ruins of an originally
elevated installation of 1.2m x 2m). The bellows supplying it with air must have been
located, according to several small post-holes, at its edge. A secondary fireplace was found
the SE corner (diameter about 50cm). It was probably used to heat a stock of iron bars
prior to their working in the main hearth. A small pit in front of it represent the traces
of an anvil bed. A compartment full of earthy substance used instead of sand as flux in
welding operations, adjoined the hearth (Bailly-Maı̂tre 1995, 338, fig. 3: B: B 68). A
channel supplied the forge with running water. At the eastern wall, near the entrance,
a depression was filled with slag incl. the PCB cakes (about 12cm in dia., Fig. 63: 2)
By the way, 3 tonnes of slag (chemically analysed by A. Ploquin, o.c., 341 - 345) was
deposited in the surrounding area having been thrown away from another, unidentified
smithy.
B 86 provided better evidence than the more incomplete case of B 71b, the other smithy
on the platform. It was partly excavated, the eastern long side was found to be absent.
Nonetheless, there was a squarish hearth substructure (150cm x 200cm) nearby, at the
western wall of the 4m x 7m room. The bellows position might have been 1m to the north
(stone pavement and some post-holes), flowed around by a similar channel as in B 68.
The site of Gua is located in a 2km away from the village. The local smithy was a
small room (2m x 3.5m of internal and 4.14m x 3. 8m of external space) housing an
originally elevated hearth (about 60cm x 60cm, of a similar type as in B 68 and B 71b).
This feature was in the worse state of preservation.
Smithy B 25, on the contrary, provided much information. It was situated close to
numerous adits at Lac Blanche at a height of 2700m above the sea level. Intensive mining
180 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 63: Smithies at mines. 1 Smithy at a mine after Biringuccio (AD 1540); 2 Brandes,
France, smithy B 68, 13th century AD; 3 Pampailly, France, smithy at a silver mine, 15th
century AD; 4 Gastein-Bockhardtal, Austria, smithy at gold and silver mines, 16th century
AD, 2 After Bailly-Maı̂tre, 3 after Benoit, 4 after Cech and Wallach.
THE SMITHIES 181
operations necessitated the installation of the smithy since it was impossible to supply
the miners with functional tools from the valley somme 1000m below. The ground plan
delimited by stone-and-clay walls was rather irregular leaving the internal space around
2m x 4m containing a hearth substructure with a trapezoidal ground plan which was
supplied with air from attached bellows. A stone-lined box (120cm x 90cm) containing
the fluxing earth adjoined in the south. A rectangular secondary hearth (50cm x 50cm)
was in the SE corner. Two blacksmith’s tongs (30ccm and 56cm long) were found in the
atelier as well as two picks or chisels welded together from bands one of which was a high
carbon steel (Baily-Maı̂tre 1995, fig. 8). No PCB’s were reported from smithy B 25 but
there was a lot of amorphous slag. Storage halls of charcoal secured the supplies for the
smithies due to the lack of fuel at that altitude.


Gold was mined in the 13th - 14th centuries in several localities in Bohemia. In the
southern central part of the country it was Čelina. The stone substructure of a rectangular
hut (4m x 5m) was found near a rock-cut shaft. Some 10 metres to the west was an area
covered with charcoal and burnt-red clay fragments where was an elongated hearth and
about 10kg of PCB slags in the surrounding area came to light. An unfinished granite
block indicates a possible anvil position. Furthermore another deposit of PCB slag cakes
was located in the south-west (about 30kg). 300 nails and clamps, 5 mining picks, 1
spur and many iron fragments were collected in the area of that mine smithy (Kudrnáč
1984; 1992). Recently, the same author discovered a 15th century gold mine at Kvilda, S
Bohemia. Preliminarily, he has documented smithing activity in the form of PCB calottes
in one of which a point of a possible mining pick was still sticking (Kudrnáč, personal
communication).
One of the best examples of the high medieval mine smithies was yielded by work at
Pampailly, about 20km km eastwards from Lyon, France. This is a hill (519m) where the
river Cosne flows round the northern perimeter. There were two smithies at the silver
mine (Benoit 1997, 18, fig. 7). The western is denoted as ‘forge de Cosne’, an earlier
and a later one are known from written sources (inventories) from the second half of the
15th century (ibidem, 52 - 53), which mention sets of nails, anvils, hammers, tongs, chis-
els and bellows which represented more than 22% of the mining expenses. The second
smithy was situated 1km eastwards downstream the river from the Pampailly hill at the
site called Vernay, close at the mine adit entrance. The stone-built ground plan (Fig.
63: 3) consisted of two rooms (4.4m x 5.8m and 5.8m x 7m, the latter, north-eastern
one, having been the workshop proper). In the north-west an annex housed an oven for
heating, the second floor, a clay-and-brick construction, having served as a dwelling. The
smithy was equipped with an elevated stone hearth (destroyed, 100cm x 50cm) blown by
bellows placed on a wooden timber imprints of which were preserved, both adjoining the
NE wall. In front of the hearth, a pit surrounded by hammer scale flakes, indicated the
anvil base. The floor was covered with both charcoal and mineral coal for heating the
metal (the latter came from Saint-Foy 8km away). There was a compartment storing iron
destined for recycling at the SW wall. The Vernay forge yielded 10 iron mining picks,
sophisticated heat treated tools with welded-in steel points, heat treated (Guillot and
Fluzin 1987, 251 - 252, fig. 7; Guillot et al. 1987, 412, fig. 2; 414). According to the
preserved mine inventories and bills, 960 iron picks were in use for 50 miners working with
182 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
hammers, i.e. 19 picks which were regularly damaged and required a repair in the smithy
for one miner.


Two post-medieval (16th century) examples should be added to complete this picture of
mine smithies: first, the forge at the Samson silver mine in Alsace (near Sainte-Croix-
aux-Mines). This was a stone-walled building comprising two rooms: the smithy with
three hearths (6m x 9.5m) and a charcoal store (6.5m x 4.5m). The bases of all of the
hearths also were stone-walled. Hearth A was in the NW corner (internal dimensions
80cm x 66cm) and its surrounding area labelled as zone I yielded completely unusable
iron picks among the iron fragments. Heart B was leaned against the eastern wall and
measured 150cm x 86cm. What could have been the anvil bed was set in front of it.
Zone II in the NE corner of the room contained iron nails and a stock of worn mining
picks. The substructure of hearth C was found at the southern wall (internal dimensions
160cm x 60cm) and, according to several post-holes indicating the bellows position, was
blown from the east. This area was zone III from which 73 mining picks were recovered
Bailly-Maı̂tre 1995, 335 fig. 2, 336; Grandemange 1990).
A lack of information on the slag waste was not a problem at another mine smithy that
was excavated at Gastein-Bockharttal, Austria. At that site silver and gold were mined
during the 15th - 16th centuries in an altitude of 2100m. The smithy excavated has been
dated to the later period but traces of an earlier smithing activity are attested without
any doubt (Cech and Wallach 1998; Cech and Wallach 2000). The smithy was located
in the SE corner of a complex dry-walled building (9m x 11m, Fig. 63: 4). It measured
3.5m x 4.5m and was equipped with an elevated hearth on a stone substructure (200cm
x 100cm) placed in the SE corner of the smithy room which had a wooden floor. There
was a quadrangular stone block with a slot for fixing an iron anvil (60cm x 60cm) and
a water barrel in front of it. Another hearth (70cm x 70cm) adhered the eastern wall
and was denoted as an essay installation for testing the ores mined. The L-shaped rest of
the house was a store for charcoal and other materials (Cech and Wallach 1998, fig. 15).
The smithy destroyed a 15th century site which had included an earlier ironworking shop.
The PCB slags of that earlier period are smaller (dia. about 9cm) and were found in a
shut down shaft and in a heap westwards of the house complex. The slags of the 16th
century smithy were mostly deposited behind the eastern wall. They were, in sum, larger
(diameter 10 - 12cm). The main task of these smithies was to keep in order the miners’
picks. 20 complete items were found and several hundreds of broken points (Cech 2000,
27). According the written sources a miner (AD 1530) ‘consumed’ or damaged 10 picks a
day so that he had to wear a set on a leather string over his shoulder. All of them had to
be reforged. Apart of picks, many other kinds of iron objects were found, 6000 pieces in
total, among them about 1000 nails, horseshoes, fittings and parts of mine-tubs (‘Hund
and the like). The PCB slags were submitted to a trial to classify them according to their
shape (Cech and Wallach 1998, 120 - 121, fig. 13) . However, they were waste and not
deliberately manufactured artefact.


The blacksmiths at mines, mostly living in unfavourable conditions and facing problems
with regular fuel supply, rendered services in taking care of iron equippment needed in
THE SMITHIES 183
the miner’s work but above all their main task was to maintain, repair or even produce
iron picks, the principal miner’s tool which must have been of the best quality.
A smithy at mine is nicely illustrated in the ‘Pirotechnia’ by Biringuccio (1540, Preface
to book I: ‘buildings for working iron to mend broken tools’, see the MIT edition 1966,
17). Fig. 63: 1.
Chapter X

THE SMITH’S PRODUCTS

The progress of the blacksmith’s work in history is reflected in the number of kinds or,
rather, functional categories of artefacts. In chapters IV and VI products like bars, ingots
and smithing tools were discussed in detail. On the present pages blacksmith’s products
involving weapons, ornaments, tools, agricultural implements, domestic utensils, personal
gear as well as structural iron are treated in general as they occured à travers les âges in
European material cultures, their settlements, graves, and hoards.

Iron inlays

Before discussing this subject, it is necessary to comment on a functionally passive role


of iron: that of ornamenting bronze artefacts in form of inlays. I hesitate to count them
as an artefact category.
In Europe, this took place in the late Bronze Age (HB3 – 9th/8th centuries BC). Forged
iron wire was used for inlays on knives (Kjøldbymagle, Sweden, ca 8th century BC, Stjern-
quist 1961) and sword hilts which are known from sites in Switzerland, Germany, Poland:
Mörigen, Helpfau, Töging, Gailenkirchen, Wald a.d. Alz, Unterkrumbach, Gamów (Fig.
III 1: 3 - 5, see Kimmig 1964, 274 - 275; Pleiner 1981, 118, fig. 5: 1 to 10). Minute
inlays appear on bronze bracelets and clothes pins as well (Mörigen, Zürich-Alpenquai,
see Fig. 64: 1 - 2; Kimmig o.c. 279; Pleiner o.c., 119, fig. 6: 2 and 3). At that time iron
was a precious metal in Europe. Its ornamenting role, although when polished, seems
to be dubious in the light of a strong tradition alluding to iron as a symbol of strength,
resistance and legitimacy in terms of written data that has emerged on innummerable
cuneiform tablets of the Hittite confederation (Siegelová 1984). Therefore, rather magical
and symbolic reasons should be awarded to this.
Certain confirmation can be inferred in the case of the bull statuette from the Býčı́
skála cave in W Moravia. It belongs to the rather later period of HD (5th century BC)
and should be considered as a late survival because at that time iron was used for the
manufacturing of many objects (weapons, tools etc., and smithing tools were discoverd in
in the same cave finds complex, too). The bull was cast in bronze (height 11.4cm, length
9.7cm, Fig. 65) and was inlaid with iron triangles on both shoulders and with another one
on the forehead and an iron wire followed the spine. The bull is a formidable cast and its
iron inlays indicate that the master was acquainted with the handling with small amounts
of iron. It is said that the stauette was found in front of the cave entrance deposited in a
pottery vessel together with charred millet. However, the hoofs bear traces of riveting so
that originally the bull must have been part of a larger object, either the top of a scepter
or a vessel handle. The find circumstances are hard to verify (1880’s) The iron inlays
were not part of any ornamenting elements and fulfilled a symbolic role conforming with
ancient traditions.
The bronze bracelet of the HD period at Czȩstochowa-Raków was equipped with inlays
of high nickel-rich (meteoritic?) iron: 8% - 10% Ni (Piasowski 1970).
THE SMITH’S PRODUCTS 185

Figure 64: Iron inlays on bronze artefacts, final Bronze Age. Ornaments: 1 and 2 Zürich-
Alpenquai. Sword hilts from Germany: 3 Töging; 4 Wald a.d. Alz; 5 Unterkrummbach,
Mittelfranken. The iron inlays are marked as thick black lines. Pleiner after Rieth and
Müller-Karpe.

Iron artefacts in graves and hoards

Specific categories of iron objects appear in graves. They were very few in number
having been subjected to a strong selection process, especially in the case of male and
female burials. Leaving aside the so-called ‘smiths’ graves’ with smithing tools which were
discussed in Chapter VI, the number of categories would exceed a little more than two
tens. In earlier periods above above all weapons and possibly horse-gear use to be put into
the graves of men. A nice example is the Greek late Protogeometric incineration from the
Agora in Athens (about 1000 BC) comprising, besides pottery vessels a bent flange-hilted
sword, two lanceheads, a knife, a trunnion axehead and a bridle bit. Without any doubt
186 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 65: Býčı́ skála cave, Moravia. A statuette of a cast bronze bull with iron inlays
(dotted): a strip on the spine, triangles on shoulders and forehead. Heigth 11.4cm, length
10.7cm. 5th century BC. After Nekvasil.

it was warrior grave. Iron objects, rather prestige goods, were also later buried in the
graves of leading figures as well. Here we can consider as an example the cemetery of
Hallstatt in Austria where inhumations and incinerations were uncovered in picturesque
arrangements, iron weapons, swords and daggers, axes and knives were added to richer
graves together with artistically shaped bronzes. Tools, pincer tongs, chisels, files, spits
occured sporadically (von Sacken 1868; Kromer 1959). Chieftans’ graves in the late
Hallstatt Bylany culture in Bohemia hid, apart iron swords and knives, waggons with
nailed tyres and lynch pins.
A different group is represented by the flat inhumations of the Middle La Tène period
in Central Europe. The warrior burials containing the full armament of the time are
particularly eloquent: long iron swords, lanceheads, shield bosses and rims, helmets, in
certain cases iron punched girdle chains (in German ‘Panzerketten’). Iron objects were
absent from female graves apart from some personal ornaments. A similar composition
of grave inventories is displayed by the incinerations of the Bohemian Kobyly culture
(plus shears and tweezers) or the pit graves of the earlier Przeworsk culture in Poland
(Kostrzewski 1949, 173 - 181, pl. XVIII). Again, weapons, some utensils and personal
ornaments appear among the grave goods.
The Early Middle Ages brought some changes due to Christian customs. The grave
ceased to be equipped with expensive goods and early Medieval graves rather contained
parts of the personal gear, above all knives, according to Borkovský (1957) the sign of
a free man at least in Slavic cultures, clout nails on boots, strikes-a-light. Christianity
spread from above in central Europe, from the higher social levels and gradually permeated
THE SMITH’S PRODUCTS 187
down to the wider population. In certain regions, e.g. in Bohemia, burial mounds were
still constructed and considerable richness was put out of circulation. Men of the 10th
century AD were still buried with their armaments (swords, daggers, battle-axes, spurs)
and even other valuable things in flat inhumations. A sword (a sophistically conceived
steel weapon, metallographically examined), axe and precious objects like a gilded silver
chalice, silver belt fittings and glass vessels were found in the male part of a 10th century
AD double grave at Kolı́n in central Bohemia whereas; the wife’s grave contained filigree
ornaments and an iron-fitted cauldron.
During the High Middle Ages, before the time when the dead were buried in shrouds,
the men simply wore everyday dress in their graves completed with a girdle belt satchel
containing a knife and a steeled strike-a-fire.


The hoards already mentioned in Chapter IV (iron ingots and bars) and VI (smith’s
tools) represent, on the other hand, smaller or larger assemblages consisting merely of
other objects, various iron agricultural implements, tools of every kind, weapons, fittings,
dress- and horse-gear, often together with bronzes, vessels and ornaments.
A group of everyday artefacts left in an abandoned interior of dwellings, for example,
should not always be understood as a hoard. A hoard is an intentionally deposited assem-
blage of things. Hoards containing irons have been found in settlement areas and hillforts
as well as in remote places outside inhabited agglomerations. The finds circumstances of
many of them, especially those from the 19th century, and accidentally uncovered fea-
tures, remain unknown whilst in other cases the intentional burying was evident: the
objects were hidden in stone settings, special pits, or in vessels (both pottery and metal
cauldrons/kettles) or in wooden chests or were wrapped into a piece of cloth.
There are hundreds of such mass finds in European archaeology. The following words
are devoted to characteristics of several hoards from various periods of the European Iron
Age just to present an imagination what values in what numbers could have been hidden.
Of special interest would be the individual functional categories of iron objects.
Two La Tène period hoards (1st century BC) may illustrate some differences in their
composition. In 1997 at Bezdědovice near Blatná, S Bohemia, a hoard of iron objects
was discovered in a stone setting and covered with a stone. It contained 48 complete
artefacts and 137 fragments (total weigth over 7kg). No weapons and explicit agricultural
implements were registered but tools (incl. two sledge hammers and a forge spoon, 5 tools
were metallographically investigated), horse-gear and numerous fittings, wheel (nave)
tyres, rings, chains, locks etc. A contemporaneous settlement existed in the close vicinity
where iron artefacts also have been found (knives, wheel tyres, brooches, see Michálek
1999). The hoard of ironwork in another Bohemian locality, namely Kolı́n the central
part of the country has a slightly different character. It was discovered in 1936 and
counted 68 items (about 15kg): a large kettle hook with chains (140cm), a forge spoon, a
hammer, adzes, chisels, six socketed axeheads, a symmetrical ploughshare, three mattocks,
a shovel blade, sickles, a scythe sleeve, eleven knives, five shears, a sliding lock key, a lynch
pin, wheel tyre fragments and a bridle bit, chains and rings. Weapons were meagerly
repesented: a spearhead, two shield bosses and an iron shield rim. Unfortunately, the
find circumstances are not known - the find is supposed to have been uncovered in one
of the two factories but what was quoted was the depth - 70cm below the surface level
188 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
(Rybová and Motyková 1983) so that it is hard to speculate in terms of the reason for
burial.
Three examples from Roman hoards will be briefly mentioned. That from Brampton,
Cumberland in England (1964), with a number of damaged hoes, ploughshares, scythes,
axeheads, augers, chisels, chains which was uncovered in the area of a military station
and was interpreted as a material deposit for scrap (Manning 1966). The huge deposit
from the fortress of Inchtuthil at the Scottish border deserves special attention: 763 840
nails of various sizes were sealed in a pit to be hidden and not to be spoiled by the
attacking Scotts when the legionary fort must have been evacuated about AD 400 (Pitts
and St Joseph 1985, 105 - 111, pl. XX). Another type of late Roman hoard is being
described by Marichal (2000) from Ruscino in the eastern Pyrenees. The matter is of a
pit filled up with 113 ferous objects including knives, sickles, scythes, vintner knives, a
tripod, punches, chisels, adzes, mattocks, hammer- and axeheads, shears, tweezers, a pan,
ironbands, keys and rings, a fragment of an iron candelabrium. In addition to that bronze
sheet and bells and a pottery formed part of the content. The assemblage dates from the
early 5th century AD, i.e. from the time of the Visigothic invasion.
Early Medieval hoards have a specific nature since they mainly consist of agricultural
implements. Nonetheless, the presence of horse-gear, stirrups, attaches and mountings and
even weapons (arrowheads, spears, axes - the latter being rather an exception) should be
noted. This picture originates in numerous hoards of the Early Medieval period (7th/8th
- 8th centuries AD) from Bohemia, Moravia, and Slovakia as has been was presented by
Bartošková (1986). The assemblages which follow are selected examples. Of a pre-Great-
Moravian date (before the 9th century AD) is a small deposit from Přı́tluky, S Moravia
put in a pottery vessel: a symmetrical ploughshare, two hoes (or rather adzes) and one
axehead - that was all what was earthed (o.c., 49, fig. 15B). On the contrary, the relatively
contemporaneous hoard from the Slovakian Gajary-Pustatina Vrablicova II o.c., 13 - 16,
fig. 5 and 6) comprised about 50 items: apart of mountings and fragments 7 sickles, 8
scythes, 2 coulters, 1 mattock, 2 shears, 1 axehead, 3 iron discs or dishes or plates, 1
strike-a-fire, a sliding lock key, and, in addition, 1 spearhead, 1 bridle bit, and 3 stirrups
as representants of the gear of armed people. The two hoards from that site (I and II)
are said to belong to a neighbouring settlement but no details are reproduced. The Great
Moravian period (9th and early 10th century AD) is represented by 32 hoards. Mikulčice
(a princely site of the time) number II yielded more than 65 irons of at least fifteen kinds
or categories (from four hoards found). This is called the ‘large depot’ in comparison with
the ‘small depot’ counting about 18 items of 8 kinds. Both were placed near the feature
called a smithy (unpublished) in the acropolis near the three-nave basilica. They date
from the final phase of the site in the first years of the 10th century AD (o.c,, figs 10B
and 11A and B) when Magyar tribes endangered the Megalé Moravia.
Unfortunately, there is not possible to discuss here numerous hoards from eastern and
south-eastern Europe, the catalogue of which, 92 complexes in number (including those
from Great Moravia) was published by Curta (1997, 253 - 261) who has introduced a new
idea in terms of their interpretation (see below).


The intentional hiding of considerable amounts of things made of a valuable material
like iron and other metals has provoked discussion concerning the reason for concentrat-
THE SMITH’S PRODUCTS 189
ing these artefacts out of sight of any undesirable person. Following reasons come into
question:
First, the temporary hiding of wealth which could be later retrieved to avoid any steal-
ing or taking as booty in uncertain situations or periods of raids or wars. Second: the
burying of values not to be taken by enemy without any intention to use them again
(withdrawal of people, evacuation of troops). Third: a votive deposit as a sacrifice with
the actual rite remaining a mystery. Fourth: the depositing of iron hoards after finishing
of socio-economical (tribal) feasts described by Curta (1997) and compared to a kind of
potlatch. This was practized by individual groups of NW American Indians involving
the host’s lavish distribution of various gifts (requiring reciprocation) as an expression of
alliance of different tribal collectives. Instead of armoured conflicts the gifts mentioned
would be removed of circulation and buried. They were parts of the property of groups
manipulated by the chiefs. The latter interpretation of hoards found in south-eastern
Europe is doubtless a very interesting idea; however, it implies a danger of generalization.
Just the assemblages fron the Slavic territories can be explained in terms of the three pre-
vious possibilities. For example Novotný (1969) takes the early Slavic (and Scandinavian)
hoards as traces of a symbol reflecting the agrarian cult or worship because agricultural
implements are so strongly represented in their inventory.
Any solution could be taken into account considering the mode of burying or the plac-
ing in remote places or settled areas.

The amassed deposits of metal objects, above all iron, which may be with little doubt
viewed as sacrifices appear in a different light: that is goods thrown into lakes, rivers and
moors.
In this respect the eponymous locality of the La Tène at the eastern sandy shore of
the Lake Neuchâtel in Switzerland presents itself. Due to two discovered timbered moles
around which immense quantities of goods (iron, wood) had cumulated on the lake bed and
the find was previously held as a trading or customs post until Raddatz (1953) explained
it as a place of sacrifice where the Helvetii tribe had thrown valuables into the water as
offerings (thankgiving for victory, begging for success). Around 2500 artefacts of about
70 kinds were found: weaponry (swords - some of them punchmarked and wrapped into
cloth, lanceheads, shields, arrows), then ornaments and fittings but also human skeletons
and skulls (Vouga 1925). At Port near Nidau in the same region, on the south-western
end of the lake Biel and in the Zihl river about 60 swords, 14 daggers, 35 lanceheads,
an iron helmet and other 150 iron objects (knives, shears, chains) were found. Similar
places came to light in other parts of the Celtic world. In the lake of Llyn Cerrig Bach,
isle of Anglesey (Man) in the Irish Sea numerous swords in scabbards, iron tools and
bars, wagon mountings, chains were found sunken (Fox 1946). The island of Mona used
to be a centre of druidic cults. Celtic swords and parts of ring mail, however out of the
Celtic domaine, were discovered in a sunken boat in the moor of Hjortspring, Jutland.
Offerings, above all swords, used to be thrown into rivers as well (Chalôns-sur-Saône,
ancient Cabillonum, France). The well known La Tène period iron hoard from Kappel,
Württenberg, SW Germany (fire-dogs, tripod, scythe, knives, blacksmith’s tools, bronze
toreutics, often damaged) was explained by Fischer (1959) as an artisan’s deposit; what
190 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
is suspicious, however, is that the find was dispersed in five groupings an located in the
moor.
As a specific case can be regarded a rectangular Celtic sanctuary within the oppidum
of Gournay-sur-Aronde near Compiègne in France (Brunaux et. al. 1980; Uran 1983, 6 -
15). The surrounding moat was full of sacrified armaments (swords, lances, arrowheads,
shield bosses, many swords were investigated by Uran o.c.).
Among Germanic tribes having inhabited the western Baltic regions the custom ruled
to throw weapons into moors in selected places during the Romano-Barbarian period.
Sites like Thorsberg, Nydam, Illerup, Vimose in Jutland are famous. Thorsberg near
Schleswig (Germany), a vaste place of sacrifice (1st century BC up to the 4th century
AD), yielded swords, lances and silver things; Vimose, isle of Fünnen, Denmark, the moor
hid weaponry (2nd to 3rd century AD); Nydam in Jutland yielded a sunken planked boat
with an iron anchor which was loaded with 100 signed swords, lances, arrowheads, horse-
gear and in addition some Roman coins (4th century AD); at Illerup in eastern Jutland,
3rd - 4th centuries AD, a deposit of sacrifial offerings was buried in the moor, mainly
swords (metallographic investigation by Thomsen, 1992).
There are not known mass deposits of that kind from later centuries, presumably the
coming of Christianity influenced the pagan rites.

Iron finds from settlement layers and the scope of artefact production

It is an irony of fate that the largest amounts of artefact categories, not regarding various
isolated stray finds, come from settlements. However, iron was a valuable material and
did not used to be left out of circulation without serious reasons. Some lost items and
fragments remained on sites that had been abandoned in peace (e.g. abandoned and evac-
uated villages). A different situation arose in long existing settlement centres: hillforts,
oppida, towns, where, dramatic events might have taken place. Assaults and raids by
enemies, fires and large conflagrations caused a lot of things to be buried in the debris
and never recovered and, what is important, without any selection. These items are the
main subject of the following discussion. Some items have to be inferred from remarks in
written sources.


Iron as an originally prestigous metal came into common use from the top levels of society
but gradually it spread down to the common people. This development can be seen by
observing the situation in ancient Greece, the first European region where a fully-fledged
civilization using iron was achieved (Pleiner 1969a), here Fig. 66. It comprises more than
70 categories of artefacts, under ten items are attested in written sources (mostly heavy
objects having served in warfare). The following smith specializations are mentioned
in written sources: the making of armour (hoplopoiiké), swordsmiths (machairopoioi,
xifourgoi), cutlers or sickle-makers (drepanourgoi) and hoe-makers sminyapoioi).
Ironwork throughout the Roman world is traceable in the Imperial provinces so that
first the Celtic La Tène civilization shall be briefly introduced. Not regarding about
twelve categories of smithing tools another 75 kinds have to be mentioned, nearly 90 in
THE SMITH’S PRODUCTS 191

Figure 66: Development of the assortment of iron artefacts in ancient Greece. Black:
archaeological object, black-and-white: proved by finds and texts, white: mentioned in
written sources. After Pleiner.
192 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
total. We have to list socketed and shaft hole axeheads, shell augers, leather knives, planes
and razors amongst the tools. In the agriculture scythes with iron sleeves, shovel blades,
fish hooks found their use. Domestic utensils: fire-dogs, balances, forks, grids, needles
as well as large chained kettle hooks, sliding lock bolts enriched the inventory of iron
things. Spurs can be added to horse-gear, shield bosses and rims, and ring mail to the
armament. Diverse fittings are sometimes difficult to classify as to their function. Large
nails were used in tonnes in the timber and stone construction of muri gallici, clamps of
different kind, and waggon mountings like wheel and nave tyres and lynch pins are not
rare finds. Iron chain belts can be mentioned among the personal gear, not to mention
fibulae and ornaments. Iron fetters are interesting objects. Iron sheet was produced. The
monumental and ritual sphere is represented by the famous iron sheet horse of Manching
(Krämer 1989). Bars and ingots were discussed in chapter IV.
Roman provinces, their towns, fortresses, vici and villae rusticae are a gold mine in
terms of individual functional categories of iron artefacts. About 100 of them can be
mentioned. Smithing and metalworker’s tools include hardies and soldering irons and
amongst other tools there are nail-pullers and various shears, trowels and vices. The
agricultural assortment covered, in addition, rakes, prongs, coulters, bronze-coated cattle
bells, capacity measuring irons (Magdalensberg). Weigh beams and pairs of compasses
deserve attention as well as padlocks which were the products of locksmiths. Fittings are
difficult to attribute to individual functional groups; door hinges became an important
structural item. Girdle buckles became common. Specific spearhead (pilum) and iron
projectiles should be underlined, and armour: loricae segmentatae and ring mail. Roman
blacksmiths managed to create immense iron objects, heavy blocks or beams destined for
the construction of bath heating installations; their weigth exceeded 250kg. The special-
ization of smiths is reflected in the terms for cutlers (cultellarii), file makers (clavarii),
shoe nailers (sutor cerdo) and locksmiths (claustrarii, clavicarii).
It goes without saying that the non-Roman, Barbarian territories used a meagre num-
ber of items, especially in the 1st century AD when the production of iron began to
supply limited or wider groups of inhabitants (for an outline of the delopment of Romano-
Barbarian ironmaking see Pleiner 2000c, 45 - 48, 272 - 275). Apart of personal gear and
basic implements the assortment consisted of weapons though their abundance could have
been limited in certain regions as can be understood in the light of the famous phrase in
Tacitus’ Germania 6: ne ferrum quidem superest sicut ex genere telorum colligitur.


It is difficult to treat the conditions in the Early Middle Ages both in the former Roman
provinces and regions colonized by newcomers since the smaller number of completely
investigated and published sites can give rise to misinterpretations. A certain overview is
provided by a survey of the items hoards (see above).
The High Middle Ages provide the following picture: The number of iron artefacts
categories is, again, higher than one hundred. Underlining the most interesting of them
causes repeated difficulties. Folding pocket knives were a common personal utensil, spe-
cific hunting knives were forged, spiral augers were widely used, strikes-a-light in the
form of lyra- or buckle shape with welded-on steel edges belonged to personal equipment
and various fittings and mountings occur in innumerable quantities. Attacking weapons
included all of the current items to which sabres, halberds and pikes should be added
THE SMITH’S PRODUCTS 193
and defensive armour produced by specialists. In certain celebrated centres complete iron
harnesses with parts covering all parts of the human body and heads and breasts of horses
were created; ring mail which necessitated the minute welding and rivetting of thousands
of wire rings was an inseparable part of armour. The horse-gear was enriched with vari-
ous types of stirrup and farriers making horseshoes and horseshoe-nails worked in towns,
villages and at roads. The locksmith’s work and horloge construction began to play an
important role. The heavy forgings must have required massive hammerheads and anvils
in hammermills. They must have existed since 11th - 14th centuries AD but their orig-
inal shapes and weigths are not known until the early post-medieval period. First cast
iron objects appeared signalizing the gradual introduction of the indirect iron smelting -
gun-balls, sporadic small objects like sword pommels (Piaskowski 1991a) and, later, fined
massive armour sheets.
In principle, no serious changes can be observed in the post-medieval and early pre-
industrial material culture of mankind.
Chapter XI

METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED


TECHNOLOGIES

Metallography distinguishes the crystalline structures of metals and their behaviour


under different heating and forging conditions. This short survey recalls the characteristics
of the basic structures of iron and steel.
Austenite (in honour of W. C. Roberts-Austen, a British metallurgist) is the solid
solution of carbon in face-centred γ-iron up to 1.7 % C. When cooled slowly, it is trans-
formed to ferrite, ferrite and interstitial (‘tertiary’) cementite, and pearlite. Ferrite (after
ferrum - light ductile crystals of body-centred α-iron (0.01 - 0.03 % C). When etched,
grain boundaries are revealed indicating the grain size: 1 to 8 are the main categories
in the ASTM (American Society for Testing Materials) system; additional numbers (e.
g. 10, 11) have been introduced for very fine textures. Very coarse grains signalize slow
cooling and minimal or absent forging (e.g. in the decarburized outer surfaces of arte-
facts which had been exposed to fire in conflagration or on a cremation pyre). Smaller
grains indicate more intensive forging. Sudden impacts at lower temperatures cause crys-
tal sliding distortion marked by so-called Neumann twin bands crossing the ferrite grains.
Intensive cold hammering deforms the ferrite crystals to flat formations (the hardness
and, simultaneously, the brittleness increases). Cementite, iron carbide Fe3 C (6.67 %
C) is a hard and brittle substance which occurs as intercrystalline lamellae in carbon-poor
ferritic materials and as a component of pearlite in steels up to 0.8 % C. Excess cementite
in hypereutectoid steels form cells surrounding the pearlite grains (visible as a network
on polished and etched sections). Cementite occurs in cast irons as well.
Pearlite (so-called due to its pearly lustre) is a conglomerate of minute ferrite and
cementite plates. The ferrite plates are slightly lowered through etching so that the
cementite lamellae throw shadows. Therefore lamellar pearlite appears dark on etched
specimens. The cementite coagulates after extremely slow heating under not very high
temperatures. The result is tougher and workable but softer steel consisting of globular
pearlite. Heterogeneous mixtures of ferrite and pearlite in different proportions are typical
for archaological artefacts; estimations of the carbon content are possible with the aid of
standardized models. Ferrite and pearlite, when heated to an elevated temperature and
more rapidly cooled tend to form an acicular texture called Widmannstätten (after A.
Widmannstätten, and Austrian technologist). Subsequent annealing cases its recrystal-
lization. Ledeburite (in honour of A. Ledebur, a German metallurgist) is an eutectic
alloy of iron and carbon (4.7 % C) appearing in white malleable cast irons or in steel with
more than 1.7 % C. In iron metallography graphite is segregated as microscopic flakes
in the grey cast iron.
Metastable structures are products of a rapid cooling of austenite; they are signs of
the heat-treatment of steel. Martensite (after A. Martens, a German metallurgist) is
a metastable solid solution of carbon in steel, rapidly cooled from austenite occurring as
lenses which, when sectionned, appear as needles forming a grey acicular structure of very
hard and brittle quench-hardened (martquenched) materials. Under raised temperatures
of 200 ◦ C - 400 ◦ C dark fine pearlite spots occur in the martensitic matrix. They are
known also as troostite (after, L. J. Troost, a British metallurgist). This attests to a
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 195
certain tempering of the quenched metal. Another kind of fine pearlite may be produced
by more intensive tempering (nodular form) or milder quenching, e.g. in oil, used to be
called sorbite (after H. C. Sorby, a British metallurgist). Bainite (after E. C. Bain, an
American scientist) is a dispersed mixture of ferrite and cementite as a result of distortion
of austenite between the martensite range and about 500 ◦ C; it is stable and relatively
hard and tough below this level. Pl. I; Pl XXVI: 7.
Thus, the identified structures are keys for the evaluation of examined archaeological
specimens.


The history of metallography has been treated in seminal works like those of C. S. Smith
(1960, reprinted 1986) or R. Pusch (1976). Nowadays, about 14 000 metallographic analy-
ses of iron forgings from different European countries have been published. Generations of
metallographers and archaeometallurgists have been engaged in this research: in the early
1900’s pioneers presented the first results (e.g. Hadfield 1912, Hanemann 1913, Rupe and
Müller 1916, B. H. Neumann 1927/28, Dickmann 1928, Carpenter and Robertson 1930,
Sal’dau and Gushtina 1932). In the late 1900’s the classical period of metallographical
research on archaeological iron finds began, whether on selected objects (France-Lanord
1949, Leoni 1953, Morton 1954, E. H. Schulz 1955, Salin 1956, Coghlan 1956), or on the
large-scale taking considerable assemblages from wide territories and individual sites into
account. The first to present nearly 300 analyses of tools, weapons and utensils in early
Russia from the period of 9th to 13th century AD was B. A. Kolchin (1953) who subse-
quently analysed numerous finds from Novgorod in northern Russia (1959). He was also
the first to found a special archaeometallographic laboratory at the Archaeological Insti-
tute in Moscow where his pupils and successors produced a further 10 000 plus analyses
(L. S. Rozanova, M. M. Tolmacheva, N. N. Terekhova V. I. Zav’yalov, O. N. Bgashba).
As regards the eastern territories one must not omit the metallographic work carried out
on blades from Belarus by M. F. Gurin carried on blades from Belarus (since the 1970’s).
Sets of Lettish objects were examined by A. K. Anteins and those from Ukraine by B.
A. Shramko, D. D. Fomin, D. A. Solncev and by G. A. Voznesenkaya. J. Piaskowski
is a leading personality who has mostly investigated Polish sources (but also ones from
Bulgaria etc.) and published his results in innumerable articles and treatises (since 1950).
A specialized metallographic laboratory also started work at the Archaological Institute
in Prague in 1963; since that time it has produced more than 1000 analyses of early iron
from Bohemia, Moravia, Slovakia, Germany, France and Sweden (R. Pleiner and now his
successor J. Hošek). In Sweden names like S. Modin and J. E. Tomtlund can be read
when the metallographic analysis of early forgings is the subject of interest. Danish and
German irons have been the subject of research by R. Thomsen. Following the late R.
F. Tylecote many scholars have been and still are engaged in archaeometallography: B.
Gilmour, G. McDonnell, P. Craddock, R. M. Ehrenreich, O. Crew, C. J. Salter, J. Lang
and V. Fell. A considerable contribution is represented by more than sixty analyses made
on Irish iron finds and published by B. G. Scott (1990). In France there is a working
laboratory under P. Fluzin. Recently, investigation results have appeared from Košice, E
Slovakia, where L.’ Mihok and his collaborators have examined and published metallo-
graphic analyses from all historical periods. Selected operations the traces of which have
been observed during the metallographic investigation have been tried to be repeated by
196 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
experimental forging (Pleiner, Crew, Lyngstrøm, Biek and others).
It is impossible to list the names of all the researchers here but it has to be under-
stood that efforts are ongoing and are yielding more information on details of the art and
techniques of the European master of the black craft. The bibliography is being steadily
enriched with new contributions.


The question is how to arrange a selection and survey of the immense data in order to
illustrate the technical level of ancient and early smiths. To present them according to
individual artefact categories (up to about a hundred items) is hardly possible, amongst
other reasons because all techniques were applied in their manufacture. It is more promiss-
ing to present the technologies based on selected examples and to refer to their use in
artefact fabrication. The functional parts like the cutting-edges of blades are the most
eloquent in this respect.

Simple techniques
Some introductory words should be devoted to the simple forming of one piece of material,
be it wrought iron or a hardenable steel as was available from parts of blooms and bars.
Apart from various mountings and fittings, nails, rivets, ornaments (pins, bracelets) and
utensils, tools were forged from one single piece as well, e. g. knives (Pl. V; Pl. XI: 1 -
4). Larger objects from whose bodies minute specimens were cut out is more difficult to
classify as having been made of a ‘single piece’ although the polished sections of working
parts do not show any traces of welds. For example, ploughshare backs could have been
joined separately and the same is true of the bent parts of shaft-hole axes which clasp
the blade which was possibly made of a single piece of iron. A nice example offers an
examined axehead from Kilberg 19, Ireland (Scott 1990, 80, 82, pl. 4.2.14, here Pl. IV).
Thus, the label ‘single piece’ artefacts has to be considered as an auxiliary term denot-
ing that the specimen picture does not reveal any signs of welding. Blacksmiths produced
such things up to post-Medieval times.

Working of low carbon and heterogeneous wrought iron

In earlier periods and in less advanced cultural provinces the smiths used the poorly or
heterogeneously carburized metal of undivided blooms. This is reflected in metallographic
samples of forgings made of such a material. When observed along the cutting-edges they
demonstrate the rather inferior quality of the artefacts which became blunt in a short
time. Nonetheless, they can be even observed even as late as the Middle Ages (5 % to 20
%) of investigated finds from various regions of Europe (Pleiner 1962; 1993a; Tolmacheva,
Rozanova and Zav’yalov in Ocherki 1997). In earlier times, weapons like swords and
lanceheads were not any exception.

Selected examples:

A Celtic iron sword from a Middle LaTène cemetery in Bohemia,labelled as Třebohostice


178 was 58.3cm long (point broken, see Pleiner 1993a, 87, pl. XI). An ornamented sheet
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 197
scabbard still adhered to one edge. One specimen representing over half of the blade
cross-section was cut out some 20cm from the hilt shoulders. The slag inclusions are
dispersed and reach values of 1 - 2 in the core and 2 - 3 Jernkontoret at the cutting-edge.
The structure is principally ferritic (6 - 7 ASTM, 160 - 210 mHV) with some lamellar or
partly spheroidized pearlite (280 - 339 mHV) at the ferrite grain boundaries. The carbon
content does not exceed 0.2 % C at this point. A thin copper or bronze layer, which
was possibly a scabbard splinter appearsed along one of the cutting-edges. The results of
the chemical bulk-analysis were: 0.012 % P, with traces of Ni. The weapon was of poor
quality (Pl. II).
A lancehead from a Migration Period inhumation cemetery Klučov 1, grave 16, was
investigated in 1959 (Pleiner 1962, 119 - 120, pl. XXXVIII). Sample A was taken from
the leaf-shaped blade, sample B from the socket rim. Sample A was completely ferrite in
A (differing grain size, ASTM 2 and 4, 152 mHV). Some pearlite only appeared in the
socket (sample B, ca 0,1 % C, fine grains of 6 - 7 ASTM). The lancehead was apparently
made of a piece of soft iron (Pl. III: 1 - 4). Note: The other lance, from grave 4, was
made up of soft iron sheets (Pleiner o. c., 120, pl. XXXIX). Both were weapons of low
quality.
The site of Klučov is significant for an early medieval Slavic hillfort from the 9th
century AD which was systematically excavated by J. Kudrnáč in the 1950’s. Another
all-iron implement has been investigated, namely a symmetrical ploughshare from pit 2:
Klučov 45 (Pleiner 1962, 46, pl. XLVII). A sample from the lowel edge of the blade
has shown numerous slag inclusions in ferritic matrix (varying grain size of 1 to 2 and 4,
on the Jernkontoret scale and 202 mHV). Microscopic needles, possibly nitrides, can be
observed on one spot below the original surface. It seems that this tool must have been
exposed to a higher temperature and had not beennot submitted to further forging (Pl.
III: 5 - 7.
In the case of medieval axehead Kilberg 19, Ireland (Scott o.c.) a ‘single piece’ blade
of iron and very mild steel has been placed between the two ends of a bent iron band and
welded-together forming the shaft-hole. The structures of the blade consist of ferrite and
interstitially occuring pearlite. Pl. IV.


It has to be born in mind that a bloomery smelt regularly produced very unevenly carbur-
ized sponges or blooms so that the above discussed objects might have been forged from
divided soft parts of the primary product. Smiths who worked the blooms as a whole
produced artefacts with unevenly dispersed carbon (ferrite and ferrite-and-pearlite struc-
tures). Archaeometallurgists relatively often encounter such things - mountings, nails,
weapons, and structural iron. Some words should be devoted to the final category. An
example from an early period will be commented on.
The ashlar stones of the monumental architecture of the 6th century BC in ancient
Persia used to be joined with double-hooked iron clamps sealed with lead in rectangular
cavities. Three clamps survived in Pasargadae; one in the Mausoleum of Cyrus the
Great, the founder of the Achamenid realm, and two to the north in the ruins of the Palace
of Audience and of the Citadel (Pleiner 1967c, 375 - 379, fig. 14: 1). The clamps are
about 20cm long and their cross-sections measure about 20mm x 20mm. In Persepolis,
an Achaemenid prestige site, founded in the late 6th century BC, clamps of the same
198 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
type were found in several places (lengths from ca 4cm to 20cm, see Pleiner o. c., 376,
378 - 379, fig. 14: 2 - 3; idem 1969a, 36 - 37, figs 11 and 13: 4 to 6). The one from
the Palace of the Hundred Columns has been investigated metallographically. The cross-
section revealed numerous cracks but few slag inclusions (2 - 3 according to Jernkontoret)
and a considerable heterogenity in the carbon distribution: pure ferritic structure with a
grain size 4 - 5 ASTM) merges into a ferrite-and-pearlite zone with a local occurence of
Widmannstätten; one spot even revealed a hypereutectoid structure of lamellar pearlite
and excess cementite. The metal was relatively pure (0.06 % P, with traces of Mn). This
is a typical ‘single piece’ of a heterogeneous steel (Pl. V: 1 - 3). Note: The technology
of structural iron clamps as applied in ancient and medieval mansonry will be discussed
where relevant.
The metallography of artefacts from the category of tools or weapons has shown that
heterogeneously carburized metal was the material for their production as well. When
carbon-rich structures accumulated on the back of the objects, such tools were of minor
quality. Should it have accidentally appeared along the cutting-edge, the artefact could be
quite good. The presence of these forgings in assemblages prove that the smiths could not
safely distinguish both materials. Numerous investigations have attested that throughout
Central Germany and the Polish lowlands during the La Tène and Romano-Barbarian
periods the application of that uncontrolled material was quite widespread.

All-steel artefacts

All-steel tools already emerged in the Bronze Age. Outwith Europe, in Jordan, a blade
from Pella dates from the Middle Bronze Age II, i. e. from the 17th/16th century BC
(Smith et al. 1984). It has been uncovered in tomb 4 of the eastern cemetery and is de-
scribed as small (dimensions not given) and metallographically investigated by R. Maddin
(Pella P67-145) who was undecided whether this thin object had been accidentally and
thoroughly carburized in the hearth. The carburization is recorded as uneven but it
reached eutectoid values of 0.8 % C. What is striking is that the structure is bainitic -
the result of rapid cooling. Maddin takes the idea into account that the original blade
surface (destroyed by severe corrosion) could have been martensitic.
Europe yielded different data. A Late Bronze Age (HB3 ) curved knife has been found
by V. Furmánek at Radzovce, S Slovakia (Pleiner 1981, 120, fig. 7; Furmánek et al. 1999)
on a stone setting of an incineration burial of the local Kyjatice culture, 8th century BC.
The knife is 21.2cm long. A complete cross-section composed of two samples has been
metallographically examined (Radzovce 609, see Pleiner 1986, 327, pl. 11). The metal
was relatively pure in terms of glassy slag inclusions but there was was a longer crack in
the core of the blade. The structure consisted of lamellar pearlite (200 - 280 mHV10g)
with excess cementite cells (500mHV). Just one edge of the tool back was decarburized
to ferrite in a very thin line. Chemical bulk analysis: 0.015 % P (very low), 0.027 % Mn
(bound to the slag inclusions), 0.018 % Ni, 0.494 % Cu (very high). Alltogether, this early
knife was made of a single piece piece of a very hard steel (0.8 - 1.0 % C) and might have
served as a very good, although brittle cutting tool (Pl VI). The relatively large forging
stock is unlikely to have been accidentally carburized in the reduction zone in the smith’s
hearth. It is more plausible to interpret it it as a hard part of a divided bloom or sponge
the properties of which were recognized and utilized by the smith.
Other artefacts were made entirely of steel as well: Celtic swords (Pleiner 1993a, 146,
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 199
fig. 17: 3), hammers (idem 1962, 83, pl. XIX) and the like.
An ornamented strike-a-fire from grave 93 in the Romano-Barbarian cremation ceme-
tery at Abrahám, S Slovakia, 2nd century AD is of special interest (Pleiner 1982a, 85 - 86,
fig. 3: 4, pl. VIII). The hot sparks of burning carbon, necessary for lighting a fire, were
generated by strokes on hard steel. A metallographic investigation (Abrahám 567) was
undertaken in 1967. The tool was a personal utensil with a suspension ring formed as an
8cm long and 0.4cm thick rectangular bar or plate. It contained numerous non-metallic
inclusions (3 - 4 Jernkontoret) which were unsystematically dispersed. Etching with 2%
Nital revealed lamellar pearlite (7 - 8 ASTM, 300 - 400 mHV 30g) and a network of ex-
cess cementite with locally occuring ledeburite (1.7 % C). The surface was decarburized
(cremation burial) displaying pearlite. The chemical bulk analysis revealed 0.018 % P,
0.07 % Mn, 0.089 % Cu, 0.035 % Ni. Excellent artefact (Pl. VII: 1 - 4).
Naturally, tools, especially knives, made entirely of steel were forged in later periods
as well (about 25 %, see Ocherki 1997, 95 sq.)

Forge welding of carbon-poor iron

A great deal of investigated iron objects reveal the joining of different bands of iron
by forge welding (Pl. X). In carbon-poor material the individual welds can be recog-
nized by consistent chains of slag inclusions or beaten-in particles of hammer scale, or by
discerning parts with differing properties (e.g. phosphoric iron, indicated by Oberhoffer
etching; microchemical analyses are still rare). The joining of multiple bands can be de-
noted as piling. Using parts of various stock may indicate an important economic fact:
the recycling of scrap metal. On the other hand, piled blades (if their welds are perfect or
accompanied by only minute slag inclusions) have good properties, e.g. they are better
resistant against slight bending. The same effect could have been achieved by folding a
plate of metal sheet to yield a band intended for the making of a blade. In that case, the
welded-together layers did not differ in their composition.

Examples:

The welding-together of bands is attested in the case of one of the earliest European
iron swords, that of Singen, SW Germany, which comes from a richly equipped inhu-
mation grave of the HB3 period (9th/8th century BC, see Boll et al. 1981, 45 - 51).
The investigators drilled out four cylindrical samples from 4 spots on the blade; two of
them had decomposed but two revealed an banded structure of alternating iron and steel.
Therefore, three cross-sections were later cut outof the blade (A, B, C). The metallic ma-
trix was richly permeated with slag (9 % of volume, 5 % of weigth). The slag contained
elevated manganese (9 % Mn). The blade was welded together from 2 flat bands; each
of them was piled from numerous components of ferritic, ferritic-and-pearlitc and perlitic
material. Carbon content: 0 to 0.5 % C, average 0.2 %. The spheroidized cementite in
some Widmannstätten spots suggest a longer heating under temperature about 700 ◦ C.
A steel plate had been placed the edge at the hilt. in the edge. The sword was not of an
exceptional quality but the technology of manufacture contained attributes which occured
so frequently during the entire development of hand blacksmithing. Pl. VIII.
The knife 583 from the Romano-Barbarian cemetery at Kostolná, S. Slovakia, crema-
tion grave 62/60 (see Pleiner 1982a, 88, pl. XVIII) was welded-together from several
200 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
bands of material. The welds are marked with bent chains of slag inclusions, whilst the
matrix is more pure: 2 - 3 Jernkontoret (Kostolná 583). The main weld joins two parts
of the blade. The part protruding up to the cutting-edge was ferritic (5 - 6 ASTM) and
apparently rich in phosphorus as indicated by the Oberhoffer etching. The back part in-
volves a welded-in wire with some pearlite (0.1 % to 0.2 % C, mHV 30g 250, fine grains (6
- 7 ASTM). Chemical bulk analysis: 0.275 % P (elevated), 0.014 % Cu, 0.065 % Ni, traces
of Mn. It is questionable whether the smiths deliberately applied the harder phosphoric
kind of iron for the cutting edge (180 to 220 mHV). Anyway, the resulting knife was not
of any first-rate quality.
An example of piling iron bands is represented by an early medieval pick or stonecutter’s
chisel Čáslav-Hrádek 21. It was found in layers of a stronghold and later a medieval
town in E Bohemia (11th century AD). Many irons were uncovered over the area and
seven of them were metallographically investigated (Pleiner 1962, 155 - 156, 158 - 163,
pl. XLV: 2 - 3, LIII, LVI, LVIII, LIX, LX, LXI). Knives, shears, hammers were made by
sophisticated techniques but specimen 21 was of very bad quality. It was a double-pointed
pick, 17cm long. There was an oval punched shaft-hole in its centre. Two samples were
taken from both points: slag foils divided them into four main strips and the ends had
been frayed into a fan-like form in the course of use. The structure of metal was ferrite
(grain size ASTM 2 - 3 with finer grains along the slag foils, Pl. X).
The quality of the artefact depended on the available metal and on the experience of
the manufacturer’s ability to recognize different kinds. Low quality ferrous products oc-
curred throughout the entire history of the blacksmit’s craft but their frequency decreased
as time went on.

Advanced techniques
Since very early times certain smiths have tried to make the critical parts of tools more ef-
fective by increasing the hardness of cutting-lines and points. Secondary steeling through
cementation and the combination of wrought iron and hardenable steel are understood as
advanced or sophisticated techniques which required a deeper knowledge and experience
of handling ferrous materials.

Additional carburizing

There were masters who have recognized that edges became harder when slowly heated
in their charcoal-fed hearths and kept outside the air-flow from the tuyere nozzle. In fact,
saturation of carbon took place in tools’ surfaces or cutting-edges. Nowadays, this opera-
tion is known as cementation or (secondary) carburization. The effectivity that could be
achieved depended on the process conditions, i.e. the temperature (above 900 ◦ C) and
time during which the metal was exposed to the carburization, what effectivity could be
reached. It took a period of hours to achieve a 1mm thick cemented layer. The process
could be catalyzed by admixtures of nitric substances in charred horn, bones, hoof filings
and the like. In certain cases the carbon content may even reach eutectoid values. The
object could have been carburized on all surfaces (possibly also in crucibles filled with
charcoal) or, more frequently, just along the cutting-edge: this enabled the retention of
toughness in the construction part so that some masters preferred it and protected the
back with clay. The working properties of such artefacts were, therefore, extremely good.
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 201
The carbon-rich parts could have been, moreover, further hardened by heat treatment
(quenching and tempering). It is not clear when the secondary carburization was intro-
duced but it has definitely been in Europe since the Early Iron Age.

Examples:

At Lovosice, N. Bohemia, the large timbered princely grave III was excavated in 1956
(proto-Celtic Bylany culture, HC period, 6th century BC). The chamber measured some
5m x 3m and contained a male skeleton, bronze ornaments, horse-gear, 23 pottery vessels,
a bronze-nailed yoke and a pottery plate with a moon-shaped symbol (Mondidol) covered
with charred wood. A flange-hilted iron sword and a knife were thoroughly corroded
whilst another curved knife was excellently preserved and metallographically examined as
Lovosice 62 (Pleiner 1962, 51 - 52, pl. X). It was 20.4cm long. Two samples covered all
of the cross-section of the blade. The principal structure was ferrite (grain size 4 ASTM).
Pearlite-and-ferrite bordering both sides towards the cutting-edgel a Widmannstätten
texture of 0.35 % to 0.4 % C appeared on one and a pearlite and ferrite network on the
other (0.6 % C). Chemical bulk analysis: P in traces, 0.72 % Cu. A piece of a very pure
iron had been used in the making this nice knife. After formation the both sides of the
cutting-edge underwent effective carburization. Pl. XI: 1 - 4.
Another example of an all-steel tool comes from the layer of the Celtic oppidum at
Stradonice, Central Bohemia (1st century BC). A fox-tail saw Stradonice 43 was welded-
together from two bands of mild steel (approximately 0.25 % C, the weld is marked by a
slag foil; fine ferrite-and-pearlite, grains 7 - 8 ASTM, 237 to 297 mHV,) but its cutting-
edge with teeth must have been additionally carburized to 0.6 - 0.7 % C and, moreover,
slightly quenched - the teeth are of fine martensite (Pleiner 1962, 85, pl. XXII: 4 - 6,
here Pl. VII: 5 - 7). The presence of these technological operations (cementation and
heat treatment) indicate that the above all-steel saw belongs to the category of advanced
technologies.
The ‘Ringgriffmesser’ type knife Hostýn 503 also comes from a Celtic oppidum (1st
century BC) which flourished in E Moravia (Pleiner 1982c, 96, 127, 143, fig. 4: 16,
pl. 29). Samples A and B cover the entire blade cross-section. It is an example of
combined technology. The knife was piled from several bands of ferritic iron with varying
phosphorus content (250 - 280 mHV 30g), bordered with slag inclusion chains (impurity
3 - 4 Jernkontoret). The ferritic bands proper have a higher purity (1 - 2 Jernkontoret).
What is interesting is that one edge of the back had been stiffened with a welded-on
hard steel wire, showing a ‘white line’ seam and a light carbon diffusion into the back
body. It is martensitic, 650 - 680 mHV. One band which extended to the cutting-edge
had been secondarily carburized (martensite, 735 mHV) and joined to another which was
carbon-poor. Chemical bulk analysis: 0.051 % P,, 0.046 % Mn, 0.146 % Ni. The knife
was thoroughly martquenched and this technology resulted in a very good artefact. Pl.
XII.
The next example is dated to the Romano-Barbarian period. A smaller knife from
Sládkovičovo 555 offers a clear example of intensive secondary carburizing of the blade
(Pleiner 1982a, 84, fig. 5: 1, pl. I). A pair of shears, a razor and 3 fibulae were deposited
together with it in cremation grave A. Samples A and B cover the complete cross-section.
The slag inclusions are dispersed and their amount can be estimated as 2 - 3, locally 1 -
2 Jernkontoret. In the blade back a fine ferrite was visible (7 - 8 ASTM, 137 - 170 mHV
202 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
30g), with locally occurring needles, possibly nitrides. A very low percentage of pearlite
cppears towards the cutting-edge. Widmannstätten texture could be observed on the
side. A pearlite (200 - 334 mHV) and ferrite network dominate the cuttin-edge proper.
The margins of both specimens show coarse-grown ferritic grains as an effect of surface
decarburization in the heat of the cremation pyre. Chemical bulk analysis: 0.012 % P,,
0.06 % Mn, and traces of Cu and Ni. The knife had been carburized in its cutting-edge;
during a longer forging the carbon distribution was slightly homogenized. It is a good
blade but the quality suffered in the cremation fire. Pl. XIII.
Because of its considerable time consumption the application of secondary carburiza-
tion declined in the early and High Middle Ages. Despite this fact we can direct attention
to an early medieval spearhead from Hillsborough, Ireland (Scott 1990, 123, 126, pl.
5.3.16) which is said to date from the period of the 6th to 10th century AD. A cross-section
covered half of the leaf-shaped blade. The spearhead is ferritic with coarse grains (1 - 2
ASTM). A distinct gradient of carbon content is visible on one side of the sample which
becomes steep towards the edge. Fine grained ferrite with interstitial pearlite (grain size
ASTM 5 - 7, mHV 50g 124 - 141 for ferrite and ca 400 for pearlite) was transformed to
hypereutectoid carbon content with cementite needles and, finally, to tempered marten-
site (645 mHV). The spearhead was an effective weapon: it had been heavily hardened by
deep secondary carburization and and quenched and tempered at the end only (Pl. XI:
5 - 7).
It is difficult to trace the frequency of the occurrence of additional carburizing since
more numerous artefact analyses come from eastern parts of Europe. In the eastern
Romano-Barbarian and early Slav cultures the proportion varies around 5 % while in
a little later Slav cultures it is about 11 %. In Bohemian centres of the 10th to 12th
centuries no secondary carburization could have been attested until noow. The situation
in western and southern Europe remains veiled.

Forge welding of iron and hardenable steel

What is under discussion here is the joining of carbon-poor iron (ca 0.02 % to ca 0.25 %
C) with medium or hard steel (> 0.30/0.35 % C) by heating in a hearth and pressing it
together by hammering it into artefacts. As explained in Chapter V, this was an difficult
task which required skill and experience of the master, since the properties of both kinds
od the material differed. They needed uneven optimal temperatures (lower in the case
of steel because of danger of its decarburization). Furthermore, the operation had to be
performed quickly, for the same reason, and carefully to avoid the intrusion of slag and
minute hammer scale splinters into the seam. On the other hand, the mastering of the
technique opened the prospect of utilizing various schemes of combination of both kinds
of metal in the artefact construction. Note: Sometimes sets of artefacts from important
centers were investigated and the results show that different techniques were applied also
as regards various combinations of iron and hard steel (e.g. Wilthew 1987 for medieval
London, Pleiner 1991a for Prague, Pleiner 2003 for Sezimovo Ústı́). The ware came from
individual smithies whose masters preferred their own styles in the making of tools.

Plating with steel

This was the simplest way of combining both kinds of material. When the steel band
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 203
reached the cutting-edge, a quite sharp and tough tool would be prepared for use Fig.26:
9 - 14).

Examples:

The weapon from Holubice 606, Moravia is presented here as an example of the in-
vestigated Celtic Middle La Tène swords (Pleiner 1993a, 95 - 96, fig 17: 7, pl. XXVIII
- XXIX). It was found in inhumation grave 63, is 60.7cm long and has a central rib.
Samples A 1 and A 2 cover both cutting-edges of the blade (taken about 30cm from the
point). The amount of slag impurities varied from 2 - 3 to 3 - 4 of the Jernkontoret scale.
Sample A 2: ferrite with intercrystalline pearlite, unevenly distributed (microhardness
150 - 200 mHV 300g, grain size 9 - 10 ASTM, tending to form a Widmannstätten texture.
Sample A 1 was richer in carbon and revealed lamellar pearlite (grains ASTM 6) with a
ferrite network (0.7 - 0.8 % C, mHV 250 - 300) decreasing towards the edge. Chemical
bulk analysis: 0.018 % P, 0.016 % Mn, 0.09 % Ni. Traces of Cu. The most likely ex-
planation of the manufacture method is that two unevenly carburized bars (one of them
more intensively) were welded-together. At least one of tf the edges was of relatively good
quality. Pl. XIV.
To the Early Medieval period (roughly 6th to 10th centuries AD) is dated The till-
ing implement labelled as White Fort 55, Ireland (Scott 1990, 136 - 138, pl. 5.3.28)
has been dated to the Early Medieval period (roughly 6th to 10th centuries AD). It is
described as a plougshare but, in fact, we are dealing with a coulter, another part of a
developed plough, which was placed in front of the ploughshare to cut the tilled soil. The
massive back of the implement is ferritic (170 mHV) and contains much slag in large in-
clusions. An imperfect weld joints it to a steely, heterogeneously carburized cutting-edge
(ferrite-and-pearlite, Widmannstätten, pearlite 270 mHV). The construction was sound,
the workmanship worse. In addition, the steel had worn out so that the cutting line was
of iron in the preserved state. The implement declined in quality through use Pl. XV: 5
- 8.

Steel shells

Hard steel shells backed the supporting ductile core and kept the cutting-line sharp (Pl.
XVI). This system which consumed a lot of steel was effective up to the moment when
this line was ground out. Archaeometallography has recorded this construction in the
manufacture of weapons (daggers, swords) and knives.

Examples:

A flange-hilted dagger from Cyprus, now in the British Museum, dated to the 11th cen-
tury BC has been examined (Lang 1991, figs 1 to 4). The sample covered half of the blade
which was 29cm long. Slag inclusions were not very numerous and did not suggest any to
any system. The core of ferrite (88 mHV 200g) was encased in V-shaped plates (coarse
lamellar pearlite, 200 mHV, along the cutting-edge 196 mHV). The weld, V-shaped as
well, shows a thin ‘white line’ beyond which an interstitial pearlite diffusion is visible.
The question of possibility of eventual previous heat treatment is left open by the author.
This dirk or dagger from Cyprus represents one of the earliest clear pieces of evidence of
204 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
iron and steel welding (Pl. XVI).
The hard covering of a softer blade core can be seen in the construction of Celtic
swords. A specific example has to be pointed out. A Middle La Tène sword Jenišův
Újezd 586 was equipped with side shells which were not equipped with steel but of
harder phosphoric iron (Pleiner 1993a, 82 - 83, 141 fig 14: 6, pl. III). The find comes
from inhumation 86 of a cemetery in N Bohemia. The sword blade (with a central rib)
was 49cm long but the poit was broken and absent; a cross-section has been prepared at
the break. The slag inclusions indicate three zones the central of which was very impure
(4 - 5 Jernkontoret) whilst both outer plates were much purer (2 - 3 Jernkontoret). The
metal of the sword was ferritic. The central bar grains were fine (6 - 7 ASTM, mHV 30g
150), the shells were coarse-grained (1 - 3 ASTM) and harder (200 mHV), apparently due
to the elevated phosphorus content. The grains of one of the outer shells show Neumann
bands, the other a small amount of interstitial pearlite. Chemical bulk analysis: 0.76 %
P (highly elevated), 0.014 % Mn, 0.051 % Ni. It seems that the swordsmith had used the
hard phosphoric shell intentionally (possibly no carbon steel was available). The intention
was right but the result, in terms of final quality, rather meagre. Pl. XVII.
The steel shell system survied on a smaller scale up to the Early Middle Ages in the
manufacture of knives but did not occur very frequently.

Iron-steel-iron ‘sandwich’

The conception of this system is just reverse: in the three-layer variety the steel band is
placed in the centre and the side bands are of softer iron. This solution was literally inge-
nious. The blades made in this way were sharp along their cutting-edges being supported
by ductile side bands until they were completely worn out (Fig. 26: 9; Fig. 67).

Figure 67: Drilling with early medieval augers. Investigated specimens Nejdek 149 and
Mikulčice 167, Moravia, 9th century AD. Three-layer ‘sandwich’ system.

It is difficult to guess where and when this technique started to be applied. Some
Scythian arrowheads were equiped with a steel centre as was the case with the Hallstatt
period chisel from Chojno near Rawicz in Poland. Certain Celtic swords were constructed
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 205
in that way (Münsingen, Gáta), as well as a knife from the oppidum of Alésia in France
(Pleiner et al. 2003, 119, figs 34 - 36). It seems that the Roman masters did not got to
like this method in the Continental provinces. Implications of the use of the technique
can be found out in Britain (Tylecote 1986, 174 - 175, fig. 117) but it would be pre-
mature to seek for a possible Celtic tradition. In reality, this technique became frequent
in the cutlery of the second half of the 1st millennium AD, especially in the North and
North-East of Europe. Apart from some arrowheads and adze- or hoe-like implements
the system turned out to be an ideal one in the manufacture of knives.

Examples:

Knife 11355 of the Vendel period centre at Helgö in central Sweden (6th - 7th centuries
AD) may be considered as a classic blade made by the three-layer ‘sandwich’ method
(Modin and Pleiner, 1978, 102 - 103, figs 75 to 82). A cross-section has been prepared
from the blade (B) and tang (T). The knife is 10.5cm long (blade 4.5cm, tang 6cm).
Silicate inclusions are large but not numerous (1 - 2, 2 - 3 Jernkontoret). They do not
mark the welds which were perfectly executed. After etching with Nital three zones ap-
peared: the central one revealed fine martensite (1000 to 1400 mHV 30g), bordered with
‘white lines’ in the seams (700 mHV), possibly with segregations of As (?). Slight carbon
diffusion penetrated across the welding seems. The outer bands were ferritic with traces
of pearlite in the grain boundaries. The central steel plate did not protrude to the tang.
The sides of iron were joined there an S-shaped weld which appeared in the centre (fine
ferrite-and-pearlite, ASTM 11 - 12, 180 - 280 mHV). The knife represents a perfect piece
of cutlery. The blade was quenched. Pl. XVIII: 1 - 2.
Several tens of three-layer ‘sandwich’ knives were examined from the sites of early
Russia (e.g. Kolchin 1953, 74 - 75, figs. 35: 4, 36: 3, 37) and Belarus (14 speciemens
from 8 sites examined by Gurin in 1984, for cross-sections see 313 - 314, figs 2 and 3).
The example presented here comes from a cemetery and is labelled as Glazov 4. A heat
treated central steel band is backed by ferritic iron side bands. The welding seams are
free from inclusions and are bordered by thin ‘white lines’ (Pl. XIX: 1 - 4). Another
typical early medieval exemple has been found at Slobodka, Belarus (Gurin 2001b, 149,
fig. 2: 1, here Pl. XVIII: 3). From the same country is being depicted the cross-section
of a knife found at Kletsk (ibidem, fig. 2: 3, here Pl.XVIII: 4) the construction of which
may be held for a kind of ‘fife-layer sandwich’ with additional steel side bands.
It is not out of interest that exactly the same technique was still practiced in central
Sweden in the late 20th century. The Mora-knives were in daily use by farmers who wore
them in scabbards attached to a knob. The back of such a blade clearly revealed a dark
central steel band which protruded, between the iron sides, up to the cutting-line which
was of a dark tone as well. The author of this book owned several Mora-knives, used
them and demostrated them but never had heart to cut and analyse them.


It is unclear whether complicated tools like axes should be classified as ‘sandwich made’.
The case of the Viking Age Haithabu 1, N Germany (Thomsen 197131 - 40, figs 1 to
5) shows an axehead (16.5cm long) the blade of which (iron-piled steel-iron) is encased
between the ends of a band forming the shaft-hole. The whole is reminiscent of a 5-layer
206 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
sandwich, the inserted blade that of three main zones. The centre of the blade conceals a
hard steel lamella (0.95 % C, hardnes HV 1kg159) which did not reach the cutting-edge.
It was covered by several plates of softer steel (0.3 % C, 0.5 % C, 132, 185 HV). This
bundle is placed between two ferritic iron bands (elevated phosphorus content, 0.14 % P,
121 HV). The blade was placed between the ends of a U-shaped steel band (0.5 - 0.65 %
C, 199 HV). The laps do not reach the edge-line where an overlapping of one of the blade
layers has to be mentioned. Thomsen presupposed that the blade (up to 4cm) must have
been hardened in oil or fat. No marteniste or fine pearlite occured, however, merely a fine-
grained pearlitic-and-ferritic structure. It seems that the resulting quality of this axehead
did not correspond with the efforts of the smith who had planned an excellent tool. Pl.
XX. An interesting case of failure is a blade from Szeligi, Poland (6th/7th century AD)
where tye central steel plate runs out at the blade side leaving the cutting-edge of soft
iron (Piaskowski 1966d).
The construction of an early medieval axehead from Ireland was much simpler - Bal-
lynahinch 51 (Scott 1990, 132 - 133,pl. 5.3.24, fig. 5.3.2). Again, two side plates (large
ferrite grains, 2 - 3 ASTM, 188 mHV) embraced a small steel band (spheroidized pearlite
with some ferrite, 281 - 387 mHV) and the edge had been welded-together (inclusions
mark the welding seams). The axehead was made of a piece of low carbon metal which
was folded on to form the socket. A plate of high carbon metal had been placed into the
centre and the components were welded into a blade. Carbon diffusion penetrated across
both seams (Pl. XXI: 1 - 4).


It is perhaps unreasonable to use the term ‘sandwich technique’, so typical for tool mak-
ing, in case of the manufacture of structural iron. However, the term sounded in the
article by Varoufakis on the metallographic examination of iron clamps and dowels from
the Classical Greek architecture: Parthenon and Erechthion in Athens (Varoufakis 1992,
5). The clamps are 30cm and more in length and remind us of a letter H with much pro-
longed cross-piece. In fact, they were welded together in the middle from two T-shaped
bars (Pl. V: 4). What is striking is that the smiths deliberately applied low carbon iron
and hard steel in their construction and welded them together in two ways: iron-steel-iron
and steel-iron-steel (Varoufakis o. c., 4 - 5, 7 - 9, figs.9, 11 and 12) as has been revealed
by complete longitudinal sections.The carbon content of the ferritic iron bar might have
reached 0.2 % C in places. Steel bars contained up to 5.8 % C and their structure was
pearlite with some ferrite. Carbon diffusion developed across the welds during the forg-
ing. Widmannstätten texture was observed in some of the sectionned clamps. Numerous
chemical analyses have indicated that the iron and steel used in the manufacture were
significantly pure. Varoufakis assumes (o. c. 14 - 15) that the source might have been in
the area around Laurion. The clamps used to be put horizontally into carefully cut out
cavities in collateral ashlars, the dowels were positioned vertically joining the lower and
upper stones. All cavities were sealed with lead (see Livadefs 156, reproduced by Pleiner
1969a, fig. 9).

Welding-in the steel

Smaller bands of hardenable steel were used in order to economize the consumption of this
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 207
valuable material. As cutting-edges they were either insterted between two joint bands
(Fig. 26: 12; Fig. 68.) into the back which was folded over like a groove prior to the
welding process. The technique appeared in early medieval cutlery (Fig. 26: 12). Other
tools, e.g. the miner’s picks were equipped with welded-in point as well.

Figure 68: Reconstruction of the manufacture of an Avar sabre. Holiare 102, Slovakia,
7th century AD. Inserting of a steel band (0.7 % C) into the cutting-edge.

Examples:

A peculiar case is represented by the socketed axehead from Kjula, Sweden, 1st to 4th
century AD (dating per analogiam, see Hermelin et al. 1979). Examination by Tholander
208 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
and Blomgren (o.c.) has shown that the socket (ferrite and pearlite in Widmannstätten
formation) embraced an inserted blade. This part of the tool has to be noted for its
properties. It was made by welding of two materials: steel with 0.4 % C and iron up to
0.2 % C. This blank was submitted to multiple folding and subsequent welding. The steel
component was substantially enriched in nickel which segregated into streaks along the
welds and reached values of up to 5 % Ni. The structure was martensite along the blade
line and pearlite-and-ferrite above the edge. Only the cutting-line had been heat treated.
The authors presume that the metal with elevated nickel content is not of meteoritic origin
but the result of smelting a nickel-bearing iron ore (e.g. with admixtures of garnierite).
The axehead from Kjula has to be classified as an artefact manufactured by a skilled
smith (Pl. XXII). The investigation of that object drew attention to heavy segregations
of nickel on surfaces of the heated nickel-containing forgings and their embedding in the
welding seams. This has been observed many times in recent years when microanalyses
were applied in the investigation process.
The cross-section of an early medieval knife from Menka, Belarus, shows this system
very nicely (Gurin 2001b, 149, fig. 2: 2). Unfortunately, the publication does not refer to
metallographic data (here Pl. XXI: 5).
The inserting of a steel edge was practiced in the case of a symmetrical ploughshare
from an early medieval 9th century AD hoard in Moravia - Ivanovice 137. (Pleiner
1967d, 124, pl. XXVIII: 5 - 9) One of the cut-out samples (A) reveals a picture which
could evoke the impression of a ‘sandwich’ which is unthinkable in a ploughshare. By
the way the implement was sampled on both edge sides and B was made completely of
ferritic iron (ASTM 6). Sample A: a central steel lamella encased between two parts of
the ploughshare body showed martensite, sorbite and pearlite with ferrite traces which
indicates the partial heat treatment of the edge which was possibly slightly turned aside
to be exposed to soil resistance. A very good tilling tool (Pl. XIX: 5 - 7).
The inserting of a steel point into the body of a mining pick (Pampailly V-85-3,
France), 15th century AD, must have been manufactured in another way than the French
idiom expresses (soudure en ‘gueule de loup’) which means that the steel particle must
have been put into a punched depression in the tool body (Guillot et al. 1987, 414, figs
10 to 13; Guillot et al. 1995, 510 - 511, 519 - 520, figs. 5 to 7). The pick with a punched
shaft-hole is about 12cm long, the inserted steel point measures 3cm. The body is ferritic
and the point pearlitic and bainitic (hardness 690). Accicular ferrite was formed beyond
the welds in the carbon diffusion zones. The tool does not show any traces of use. It
was brand-new or repaired, just delivered from the smithy (Pl. XXII). The smiths at the
mines must have been specialized masters.

Scarf welding-on of steel

This method of providing iron implements with steel working parts was relatively speak-
ing not very complicated. It consisted of splicing the steel components to the side of
the construction part in the place of the intended cutting-edge so that they overlapped
the rim in an assymmetrical way (Fig. 26: 11; Fig. 69; Fig. 71: 2). After sponaneous
etching during the use a broader dark strip appeared on one side whilst it was narrower
on the other. The tools manufactured in that way remained tough and ductile but their
working edges were hard and sharp (Fig. 70: left) and could be heat-treated.
The technique used to be applied to advantage in making e.g. axeheads but it was
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 209

Figure 69: Technology of the manufacture of early Slavic battle-axes. I Overlapping of


the shaft-hole, scarf-welding of the cutting-edge. II Punching the shaft-hole and similarly
welded-on cutting-edge. Based on Great Moravian samples, 9th century AD.

frequently used in the construction of ploughshares, chisels, adzes, sickles, picks, augers
and, naturally, knives. Celtic swords were scarf-welded from several bands with varying
carbon content, too. The technique discussed above was applied throughout the historical
periods.

Selected examples (chronologically):

One of the earliest tools made in this style is the rectangularly socketed 7,2cm long
axehead from the Celtic oppidum of Manching, Bavaria. The socket is closed and the rim
stffened with a narrow sleeve (Jacobi 1974, pl. 16: 282). An examination has been carried
out on a sample from the cutting-edge (Manching 715). The body was of ferritic iron
(grain size 4 to 7 ASTM, microhardness around 200 mHV 30g). A breach or slit marks a
badly performed weld on the edge part. This consisted of two plates. The one adhering
to the slit revealed crooked layers of ferrite with varying phosphorus content. The steel
cutting-edge proper was welded to this inset: pearlite with ferrite network (around 400
mHV) merging into what used to be called troostite (a kind of fine pearlite) and, then to
martensite (1000 to 1200 mHV). Chemical bulk analysis: 0.399 % P (elevated, presumably
concentrated in the inset), 0.04 % Mn. It follows that the axehead was welded from three
components - the body, the layered or twisted inset (Pl. XXIV), and the steel edge. The
tool had only been heat treated along the cutting-line, briefly quenched and emerged; the
heat of the implement caused a spontaneous tempering of the transition zone. Despite
210 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 70: Axe and adze. Mikulčice 158, based on sample 158 (left) and Tvarožná Lhota,
sample 95 (right).

the imperfect weld on the body, axehead 715 ranks with very good artefacts (Pleiner, sine
anno, unpublished).
The application of scarf welding was quite common in the early cutlery manufacture.
Here is presented a knife blade from Migration Period hut 43 at Březno, NW. Bohemia
(Pleiner 1988 [1991b], 606 - 609, fig. 6). The polished and etched cross-section (Březno
619) shows three zones marked by scarf welds. The cutting zone was pearlitic-and-ferritic
with a transition to sorbite (400 - 500 mHV). The body was also pearlitic-and-ferritic with
the local appearance of Widmannstätten (300 - 400 mHV). Some glassy inclusions near
the cutting-edge had broken as the forging temperature had decreased. The third zone is
a ferritic iron wire in one of the back edges. The finished blade was carefully tempered in
its cutting-line and the knife can be classed as a well-made artefact (Pl. XXV).
The battle axe Mikulčice 175 (bradatice, chin-shaped axehead) was found in the Great
Moravian centre of Mikulčice, S. Moravia (9th century AD). The metallographic examina-
tion (Pleiner 1967d, 84, 129, fig. 3: 14, pl. XLVI) revealed that the whole cutting part has
been scarfly joined to an iron body with punched shaft-hole. This critical part consisted
of two welded-together metal sheets: a ferritic (grains ASTM 3) and a steely, apparently
secondary carburized plate (pearlite-and-ferrite, partly Widmannstätten, transformed to
sorbite, 313 to 363 mHV). Chemical bulk analysis: 0.10 % P, 0.15 % Ni. The axehead
was intentionally equipped with steel, edge tempered: a first-rate weapon.
The scarf welding-on of steel was also not absent from the manufacture of some
ploughshares. The one from the Ivanovice hoard in Moravia (Ivanovice 136, Pleiner
1967d, 124, pl. XXVIII) was 19.2cmlong and slightly assymetrically shaped. It was made
of ferritic iron (137 to 173 mHV). A scarf-welded steel plate (pearlite with ferrite network,
transition to tempered sorbite, 302 - 309 mHV), increased its effectivity (Pl. XV: 1 - 4).
Note: ploughshares rank amongst agricultural implements the construction and quality of
which involved simply made pieces as well as high quality tilling tools. This has already
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 211
been presented on these pages. It depended on the skill of the rural smiths and on their
acces to steel.

Should cutlery be taken into account again, an example from the 10th century AD
cannot be omitted. Here we are dealing with a fragmentary knife Budeč 677), from a
settlement within a princely hillfort, at the site of Na Týnici (Pleiner 1993b, 77, 79, fig.
4: 677, pl. 16). Three zones are revealed on the etched cross-section: the blade body is
ferritic; a scarf weld marks the join with the steel cutting-edge (fine martensite 680 - 879
mHV 30g). A diffusion of carbon across the welding seam revealed microhardness 390 -
400 mHV. The interesting part is the back of this knife; a steel strip (371 - 436 mHV,
pearlite, troostite) was butt-welded, so that after self-etching on the air a striped effect
of dark and light tones attracted the eyes. A kind of a ‘striped damast’ can possibly be
considered in the classification (see below). The cutting-edge was carefully quenched and
heat treatment affected the back of the knife as well (Pl. XXVII).
The appearance of butt-welding led to the wide application of this technique in the
manufacture of special utensils (fire-steels) and, again, in the workshops of early medieval
cutlers.

Butt-welding of steel

Butt-welding means the joining of iron rods or thicker forged wires perpendicularly to
the long axis of the artefact cross-section, ‘surface-to-surface’ as Janet Lang called it.
The hammer strokes are not directed from the side but from above. Several Celtic swords
were butt-welded: Saône river (all iron), Münsingen 24663 (the core butt-welded, plated
with shells), Llyn Cerrig Bach 4 (iron and steel rods, see Pleiner 1993a, 122 - 123, 143,
146, fig. 14: 9, fig. 17: 9, 17). It is unclear why the Celtic smith used this method.
A more advantageous application was the butt-welding of a steel to the body of a
strike-a-fire. The igniting of a spark required straight vertical beats.
As regards, the outer appearance of the object played a role in the butt-welding of
steel cutting-edges, a role played the outer look. As has been noted several times already,
atmospheric influences caused a slight self-etching effect on the blade surface. In the case
of butt-welding dark tinted steel cutting-edges appeared in equal width on both sides.

Four examples:

The early medieval strike-a-fire or fire-steel was a lyra-shaped utensil, or it was formed
like a buckle. The examined example was of the former type and was found in grave 48 of
the cemetry within the hillfort of Nitra-Lupka, Slovakia, 9th century AD (Pleiner 1967d,
120 - 121, pl. XVI: 1 - 4). It was labelled as Nitra-Lupka 107. It measured 10.6cm in its
preserved state. The ferritic body (very coarse grains, ASTM 1 - 2, 185 mHV) was per-
meated with numerous slag inclusions, whilst the striking-edge, which was butt-welded,
was much more pure. Its structure was martensite (893 - 937 mHV). Carbon diffusion was
noted across the welding seam. Chemical bulk analysis: 0.45 % P (elevated), 0.16 % Cu.
This personal utensil was of perfect quality and ready to serve any user (Pl. XXVIII: 5
- 8). Note: Buckle-shaped or stick-like fire-steels survived into the early medieval period
as well (for early Russian examples see Kolchin 1953,165, fig. 138).
Another tool will now be presented: a Viking period drawing die for pulling (non-
212 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
ferrous) wire. The tool Haithabu 1 (Naumann, 1971, 85 - 92, figs 1 to 7) was a 12cm
long bar perforated with four punched holes in a groove: diameters 5mm, 7mm, 4mm
and 6mm. A cross-cut sectioned it in the level of the first drawing-hole. Layers of iron
and steel emerged after pikrin acid etching. It was apparently butt-welded but a detailed
scheme was difficult to define. Ferrite zones (150 - 200 HV) encased those of pearlite and
troostite (200 - 600 HV), and of martensite (900 HV). The tool had been briefly quenched
after the punching of the drawing-holes (Pl. XXX).
A small knife (11cm) with a butt-welded cutting-edge came to light at the 11th century
AD hillfort of Čáslav-Hrádek in E Bohemia. This was cross-cut (Čáslav-Hrádek 23,
see Pleiner 1962, 155 - 156, pl. LIII) and revealed fairly coarse ferrite grains (ASTM 1)
in its ferritic body and martensite in its butt-welded cutting-edge. The cutting-line was
slightly decarburized, probably due to sharpening during use(Pl. XXX: 1 - 4).
The fourth example comes from Lemeshevichi, Belarus (Gurin 2001b, fig 2: 4). No
data were presented in terms of metallography were given, just a polished and etched
cross-section showing the body (two joined bands, one apparently ferritic, the other one
containing more carbon) with a nicely butt-welded cutting-edge (Pl. XXX: 5). The butt-
welded objects which sometimes included combinations of iron and phosphoric iron bands
and steely cutting-edges may be considered as top quality display artefacts.
Note: Analyses of sets of artefacts from one site and one historical period may show
that different techniques can be observed on individual samples; the artefacts were ap-
parently made in several smithies and either got to the site contemporaneously or during
several sub-phases Fig. 71: bottom).

Top techniques
The techniques which required the highest level of skill, experience and craftsmanship
from the early master smith are classed as top techniques. These involved the perfect
empirical distinguishing of different ferrous materials (iron, phosphoric iron, hardenable
steeI) and an extraordinary mastery in the carrying out of minute-scale processes (e.g.
complicated small blades) and governing work with larger pieces of material (e.g. in
making plate armour, incl. the decoration). As for the manufacturing of blades the im-
portance of the outer appearance, i.e. the toning effect, can be appreciated although this
might seem to be superfluous from the functional point of view but it was appreciated as
a sign of the prestige of the bearer.

Striped blades

These artefacts appeare among weapons and knives. Essentially, the technique of manu-
facture consisted in joining iron bands or wires by means of forge butt-welding. Individual
stripes presented themselves through differing tones visible on artefact surfaces.
For example, an early medieval lancehead from a cemetery at Lutomiersk, Poland
(Piaskowski 1959a in: Nadolski et al., see Pleiner 1962, 170, pl. LXV: 4 - 5) consisted
of an iron core (grey), bordered on both side stripes of phosphoric iron (lustre) and steel
edges (dark). Pl. XXVIII: 3 - 4.
Metallographically investigated knives with visible stripes constitute part of the corpus
of top class cutlery. 13th century AD examples have come to light in Bohemia. The
knives were found in features that were archaeologically uncovered by M. Richter within
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 213

Figure 71: Technological systems of early medieval knives. 1 Three-layer sandwich; 2 scarf
welding of steel cutting-edges; 3 butt-welding; 4 and 5 stripe damast with a phosphoric
iron band; 6 pattern-welding. Bottom: construction types as revealed in horizons of
the 10th and 11th century princely hillfort of Budeč, Bohemia. Numbers relate to the
examined specimens. After Pleiner.

a monastic service settlement at the site of Sekanka high above the Ostrov cloister on the
right bank of the Vltava river bank (central Bohemia, close to the confluence with the
Sázava river). The crafts pursued in this settlement covered the needs of the monastery
and a similar system was in operation in villages and estates ruled by other monasteries.
214 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
The blacksmith’s craft filled an important role.
A complete cross-section of a knife blade with a broken tang Sekanka 251 has been
examined (Pleiner 1984, 273, 277 - 278, figs 169: 251 and fig. 170: 2, pl. 51: 7 - 8,
pl.63: 1 - 4). Three transversal chains of inclusions divided the section of the blade into
four zones: three in the blade back and one marking the weld of the cutting-edge. A
dark one with martensite was inserted between the two ferritic stripes in the back, richly
permeated with slag (4 - 5 Jernkontoret). The cutting-edge was martensitic as well. From
the side view the martensitic zones appeared as dark and both ferritic ones as light - a
typical striped blade, thoroughly quenched. An excellent piece of work by skilled cutler
(Pl. XXXI: 1 - 5).
A full cross-cut of the second knife (Sekanka 258, see Pleiner 1984, 273, 277- 278, fig.
169: 258, fig. 170: 1, pl. 54: 8 - 10, pl. 64: 3), was examined revealing the following
scheme: Slag inclusions and a slight crack formed the weld between the ferritic back and
the cutting-edge, then came a ferritic-and-pearlitic zone beyond which was the cutting-
edge proper; this has revealed several oblique zones consisting of martensite and troostite
(the blade was apparently only tempered along its cutting-line). What is remarkable is
that from the side view of the etched blade the lower weld was wave-shaped or corrugated
or serrated which gave raise to a specific effect. This waving was preasumably achieved
by impressing a twisted steel edge part into the heated inter-layer (as shown by Thomsen
in the case of a display lance head, idem 1971, fig. 5: 2, here Pl.XXXI: 6 - 7). Serrated
welds often appear on Germanic sax blades. In Poland, knives with wave-shaped welds,
have been uncovered as well. Recently, more early medieval knives with wave-welds have
come to light from Bohemian hillforts (J. Hošek, in preparation).
Stripe welding has been secondarily denoted as ‘striped damast’ (Pleiner 1998a) to dis-
tinguish it from pattern-welded damast (see below). Note: Several Roman swords (spoils
or losses tracing Roman military expeditions) have been found on the territory of modern
Poland and mentioned as pattern-welded (in Polish miecze dziwerowane). Their cross-
sections, not accompanied by side views, rather evoke the impression of striped blades
(Sobotka, Wa̧chock, Hromówka, see Piaskowski 1965b).

Pattern-welding

By pattern-welding (in German wurmbunter Damast, Schweissdamast, in Russian uzor-


chatyy damask, in Polish dziwer) a style in manufacturing blades is meant which employs
the twisting of iron and steel rods or wires and their subsequent welding. The several
resulting rods were usually butt-welded in columns. A permanent pattern of light and
dark figures occurs after abrasive grinding and etching - in the shape of the letters V, W,
S, U and the like. It is true that the pattern could not be erased but it could change
after multiple grinding which created new planes with differing figures (up to circular ones
called ‘rosettes’).
The technique is reminiscent of Damascene steel (traded through Damascus in ancient
and later times) which was, however, crucible steel which had passed through liquid stage.
After the forging, the final dispersion of the structure particles a delicate light and dark
moiré developed - the only ressemblance or link with European pattern-welding (e.g. Lenz
1908; Harnecker 1924).
Presumably the original purpose of twisting and welding iron and steel components
was to produce long bars which would be, to certain extent, resistent against bending.
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 215
However, in due time it was just the dark and light pattern which caused the blades made
in this way to become display and prestige artefacts. Significant testimony is provided by
a letter from Theodorich the Great (final 5th century AD) to a king of the Varni tribe in
which the appearance of the sword - a royal gift - was described in buoyant words (see
Chapter XII).
Swords, scramasaxes, lanceheads and knives have come to light among archaeological
finds.

Pattern-welded swords

The true origins of this technique are unknown. Two blades among Middle La Tène
period Celtic swords two blades may be classified as pattern-welded: that from Clee-
bronn, Germany, with twisted (?) iron and steel core (Fig. 72) and from Cuvio, N Italy,
with a steel core, pattern-welded side panels and steel edges (Pleiner 1993a, 117 - 118,
fig. 12; 125 - 126, fig. 17: 12. The dating of the Cuvio sword is uncertain.

Figure 72: Early La Tène period pattern-welded sword. Cleebronn, W Germany, after
Schulz et al.

The construction of genuine pattern-welded swords was realised in two arrangements: a


pattern-welded bar might have served as a blade core, or there were plated pattern-welded
side panels on the blade core (Fig. 73); in German Furnierdamast, Deckschichtendamast.
Both models were in use without respect for chronology or geographic distribution.
Roman pattern-welded swords were, in fact, long spathae used from about AD 200
in Imperial auxiliary cavalry troops. Within the Roman territory they are known from
Rhineland, Bavaria, Switzerland and many of them have been metallographically ex-
amined (Schulz 1959; Böhne 1963; France-Lanord 1964). Both varieties are recorded
as having been equipped with steel cutting-edges and sometimes showing the punched
216 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
names of swordsmiths like CICOLLUS or COCILLUS, RICCIM or RICUI, UMCORD
(see Böhne 1963, 233). However, many finds come from lands outside the Roman Empire
having made their way to the Barbarian lands. Those from the moor deposit at Vimose,
Fyn island, Denmark, from the moor at Illerup (Thomsen 1992) or from the sunken boat
at Nydam, Danish Jutland, are of particular importance (Neumann 1928, Maryon 1960;
Thomsen o.c.). The finds from Poland have already been mentioned; the sword from
Hromówka bears a brass inlay of a figure of Mars on one side at the hilt root, and figural
ornament on the other (Piaskowski 1965b). In addition, from the cemetery of Oblin, NE
Poland, four pattern-welded sword blades were investigated (Biborski et al. 2003).
During the Migration Period the Germanic swordsmiths apparently adopted this tech-
nique and mastered it well. Nearly 300 spathae (about 80 % pattern-welded) were found
in the warrior graves in the cemetery at Altenerdig, Bavaria, the 5th to the early 7th cen-
turies AD (Ankner 1996). Swedish sites of the early medieval Vendel period have yielded
pattern-welded panelled blades from Avestra and Sanda. Metallography has revealed that
their edges were of steel with about 0.5 % C whilst the steel wires in the pattern-welded
panels contained 0.1 % to 0.9 % C. The iron components were phosphoric (Thålin 1966).
As for the examples presented on these pages we shall deal with a pattern-welded
spatha from a chieftan’s grave at Bešeňov, Slovakia, 5th century AD (Pleiner 2002a).
This is interesting because of its classic construction scheme but only one of the cutting-
edges was effective. Otherwise it was a luxurious weapon, 95cm long, the short silver and
gilded guard of which was ornated with a braid/plait motif in a niello setting. Three red
almandines hung from the lower side of the guard. Two samples covering the entire cross-
section of the blade (A and B) were investigated. The blade core as well as the butt-joined
cutting-edges were forged from a very mild carbon steel (0.1 % - 0.2 % C) with a ferritic-
and pearlitic structure (microhardness ferrite 200 mHV 30g, pearlite 230 - 300 mHV). One
cutting-line (B) was secondarily carburized and heat-treated (troostite 400 - 500 mHV,
martensite spots up to 800 mHV). Both sides of the core were plated with pattern-welded
panels, one of which was totally corroded but the opposite one was preserved and consisted
of P-rich iron wires (coarse ferrite, 2 - 3 ASTM, 170 - 230 mHV) and fine pearlitic rods.
This splendidly appearing double-edged sword, might have functioned as a one-sided hard
weapon (Pl. XXXII and XXXIII). The swordsmith must have cooperated with a jeweller.
In Romano-Barbarian communities persons of various higher ranks yearned to possess
pattern-welded swords, although their swordsmiths had no experience of welding iron and
steel or no access to harder steels. So they produced all-iron imitations with twisted sword
cores. A plastic herring-bone pattern proclaiming an ostentatious weapon was visible on
the blade surface which, in fact, was not very effective. A sword from a Romano-Barbarian
cemetery from Plotiště in NE Bohemia may serve as an example (Pleiner 1976b).
The tradition of making pattern-welded swords continued in the Early Middle Ages up
to 10th century. Some references should be quoted. Germany: Böhne and Dannheimer
1961; Timpel 1963; Bühler and Strassburger 1966; Sweden: Arbman 1937; Baltic: An-
teins 1960b; 1964; 1966; 1968; cf Fig. 74, Finland: Koch 1997; Norway: Liestøl 1961;
Netherlands: Ypey 1963; England: Tylecote 1986, 194 - 196; 1987, 275; 74). Blades
which were metallographically tested were found to be display and dangerous weapons
with heat-treated hard cutting-edges.
An example from the 10th century AD will be dealt here - the sword from a richly
equipped princely grave from Kolı́n in central Bohemia (Pleiner et al. 1956; Pleiner 1962,
184, pl. LXII). Unfortunately, the sword Kolı́n 10 only survived as a 45.5cm long frag-
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 217
ment with broken off point and pommel. In spite of that it yielded important information.
The blade surface was corroded but test polishing has shown V- and W-shaped dark and
light traces of a pattern-welded panel. A sample for metallographic analysis did not reach
the blade core. However, the examined cutting-edge was of hard steel (pearlite with ferrite
network, 462 and 249 mHV) and presumably composed of three bands. A scarf weld marks
the added hard steel rod in the cutting-edge (martensite, 740 to 830 mHV). The sword
must have been of the highest quality (Pl. XXXIV). Note: Not all of the early medieval
swords were so well made. In Bohemia the ones from Jaroměř (Z type after Petersen)
and from Libice (M type) were forged from low-carbon iron (Pleiner o. c., 165, pl. LXIII).

Apart from swords other thrusting and cutting weapons used to be pattern-welded:
matchete-like saxes and scramasaxes, termed magnae cultellae by Widukind, of the Franks,
Saxons and other Germanic peoples of the post-Migration period (7th and 8th century
AD).

Figure 73: Two basic systems in the application of pattern-welded blades. 1 - 2 massive
pattern-welded core; 3 - 5 pattern-welded side panels. In both cases butt-welded pattern
rods or bars could be applicated. 1, 3 swords; 2, 4 saxes, scramasaxes (later knives); 5
lancehead.

A study by Westphal (1984; 1991) deals with more than 90 saxes from Saxon territories
of Germany. The construction incorporates both systems: the pattern-welded panelling
(Furnier on one or both blade sides and the placing the pattern-welded rod into the mid-
dle of the blade, Massivdamast). The sax from Osnabrück had a U-shaped figure (also
pattern-welded) into the V-shaped damast. The cutting-edge part often consisted of two
or three stripes joined with serrated (wave-shaped) welds. Horstmann (in Westphal o. c.
364 - 365) investigated 6 saxes which have revealed serrated welds. The steel edges were
heat-treated (tempered). The iron back parts might have contained up to 0.5 % P.
218 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Figure 74: Examples of early medieval pattern-welded sword blades. 1 - 3 eastern Baltic,
after Anteins; 4 - 6 fragments of Carolingian swords in Swedish collections, after Arbmann.

Example:

Here we are dealing with a shorter scramasax from England. It was examined by B.
Gilmour and published by Brewer (1976, 5 - 6, figs. 9 to 12). It belongs to the Dorset
County Museum collection. The length is 35.3cm. Unfortunately, the examination was
possible on the rear (back) edge where limited polishing was allowed. Alternating layers of
hard iron (apparently phosphoric, hardness HV 20kg 260) and softer ferritic-and-pearlitic
steel (about 0.3 % C, HV 200) were revealed by etching. The cutting-edge had rusted
and could not be sampled. Pl. XXXV: 1 - 3.
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 219
Pattern-welded lanceheads

Lanceheads which may be considered as display weapons are known especially from Nordic
and Baltic lands. In many cases their sockets were inlaid with copper or silver ornaments
(Martens 2002). Their leaves were plated with pattern-welded components and a number
of pieces bear serrated edge-welds.

Figure 75: Examples of display lanceheads ornated with pattern-welded stripes and ser-
rated welds. Eastern Baltic, early medieval. Based on Anteins.

The largest number has been counted in the eastern Baltic, Estonia and Latvia. An-
teins mentions more than 200 objects (1962 a, b; 1963; here Fig. 75). They must have
been widespread in Finland too (Koch 1997, 218, fig. 7, with reference). Saxon England
yielded these artefacts as well. Tylecote (1986, 196, fig. 138, table 96) discusses about ten
of them, some having been manufactured by the Furnier system. Nonetheless, lanceheads
with pattern-welded core (Massivdamast) rank amongst the top artefacts of this kind.

Example:

Thomsen (1971c) investigated a lancehead from the Viking period centre of Haithabu-
Hedeby near Schleswig in detail. The find Haithabu 116/1966 is 40cm long (the trape-
zoidal blade 27cm).The socket of ferritic iron (200 mHV kp/mm2 ) was welded to the
blade and so were the side trunnions to the socket. The construction of the blade was
complicated. Its core consisted of two rods, each twisted from ferritic iron and pearlitic-
and-ferritic steel. Stripes with serrated welds were formed by impressing twisted rods to
220 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
this core as is nicely visible on X-ray photographs. Finally, hard steel cutting-edges were
added and the weapon was ready for grinding and polishing. Then, the cutting-edges were
tempered (martensite 855 mHV, troostite 315 mHV. Pl. XXXVI. This splendid lancehead
has been uncovered in the coastal fortification at the Schlei vik stuck into a wooden pillar.
Who did it and why? Was it the result of local skirmishes or the final episode in the life
of Hedeby town (11th century)?.
There remains the question of who was armed with display lances in the North of
Europe. Hardly the rank and file warriors, rather the chiefs and their trains, retinue or
body guards.

Pattern-welded knives

Sometimes during the 10th century AD the high technology of pattern-welding descended
from the pedestal of exclusiveness and ceased to be reserved solely for selected groups of
people and was to certain extent commercialized in the sense that it spread to cutlery,
i. e. to the manufacturing of expensive knives procurable at least to wealthy customers.
The artefacts were produced up to the 14th century.
The standard construction consisted of the following scheme: the rear back and the
body of the blade consisted of several iron and/or steel stripes the central of which was
pattern-welded (Fig. 26: 8; Fig. 71: 6; Fig. 76). An exception is the knife from
Winchester, England, 13th century (Pl. XXXV: 4). Tylecote (1986, 197, fig. 139)
depicts a cross-section showing that the entire blade body upon the steel edge was welded
from curved light and dark wires or bands. In the case of all pattern-welded knives
the etched pattern displaying light and dark figures marked the items as display ware
at the first glance. No constructional purpose can be discerned. Based on the author’s
general knowledge pattern-welded knives are only exceptionally mentioned in western
Continental Europe. Two pieces have to be registered from SW Germany: Singelfingen
and Unterregenbach, Württenberg, both dating from the 13th/14th centuries (Pleiner
1979a, 250, 253 - 254, after Horstmann and Naumann).
The domain of the distribution of pattern-welded knives extended over Bohemia,
Poland and Belarus, i. e. in Slavic or once Slavic lands, since one of the sites is in
the middle Elbe river region of Germany; another isolated find came from Novgorod, N
Russia.
The earliest well described example was uncovered at Lahovice near Prague, in grave
32 where an old man was buried in a rather large grave pit (no other objects came to
light). The classical scheme shows a ferritic back, pattern-welded middle stripe and a
steel cutting-edge with heat-treated cutting-line (martensite). The date: 10th century
AD (Pleiner 1979a, 247, 252, 357, pl. IV). Two specimens from the 10th century Slavic
site at Dessau-Mosigkau in central Germany are less clear. Knives have been found in
houses 6 and 19 (presumably late 6th or 7th century AD; the knife from hut 19 had
an iron cutting-edge) and could be interpreted as pattern-welded in the light of cross-
sections; verification by surface polishing was not possible (Pleiner 1967b). The following
remark cannot be omitted: these early finds were bound to a rural context. The cemetery
of Lahovice belonged, without any doubt, to a village and Dessau-Mosigkau was a rural
settlement as well. The knives might have been from a centre - in the Bohemian case
from Prague (?), in that of Dessau-Mosigkau from Halle (?), Magdeburg (?) or another
unknown site (?).
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 221
That is to say that later finds, from about the 11th to the 14th centuries, were ex-
pressively connected to urban production and only occasionally got into the countryside.
About 40 pattern-welded knives are known from the above mentioned regions of Europe.
As to Bohemia, seven blades were found in the little monastic town of Sekanka (13th
century), three in house 11/1, two in storage pit E, and two in feature IV (Pleiner 1982b,
277 - 278, fig. 16 and 17, pl. 44: 1 - 3; 46, 4 - 6; 51: 4 - 11; 53: 3 - 5). The knife from
Mutějovice, Bohemia, uncovered in hut 8 dates from the 12th century. The blade is made
in a classic pattern-welded construction scheme for knives. Mutějovice was a village and
the knife must have been imported from a town smithy (Pleiner 1969b, 561 - 562, fig 17
and 18: 423, here Fig. 76).

Figure 76: Reconstruction of the manufacture of an early medieval pattern-welded knife.


Based of sample 423 from Mutějovice, Bohemia, 12th century AD. 1 Presumed technology;
2 - 3 blade side, 5x, etched with Picral; 4 cross-sectioned sample etched after Oberhoffer.
After Pleiner.

Polish towns have yielded about twenty pattern-welded knives, i. e. Opole, site
‘Ostrówek’ (11th and 12th century); Cracow-Wawel (4 knives, 12th and 13th centuries);
Biskupiń (12th/13th century); Wroclaw (7 knives, 13th century, one example 14th cen-
tury), Gdańsk (13th and 14th century, see Pl. XXXV: 5). The blades were metallograph-
ically tested and revealed the usual construction schemes; the steel cutting-edges were
mostly quenched.
222 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
More knives have to be mentioned from further east: that from Vitebsk, Belarus (Gurin
2001a, 149, fig. 2: 4, here Pl.XXXVII: 9) and that from Bukhak, Ukraine, 13th century
(Voznesenskaya 1999, 113 - 114, figs 1 and 2, here Pl. XXXVII: 1 - 8), etc. According
to the cross-cut it seems that three pattern-welded rods were butt-inserted in the middle
of the blade. The knife from Novgorod Velikiy, far to the North-East, was published by
Kolchin (1953, 79 - 80, figs 40 and 41). It was uncovered in town layers of the 11th - 12th
centuries. The blade is heat-treated in its cutting-edge (martensite, o. c., 85, fig. 45: 1).

Example:

We have chosen a pattern-welded knife from Sekanka, Bohemia, dated to the 13th cen-
tury (Pleiner 1982b, 273, 277, fig. 169: 249; 278, fig. 170: 4, pl. 41, 10 - 11). Specimen
Sekanka 249 offers a survey of the complete cross-cut which reveals the fine ferritic struc-
ture in the back stripe (around 200 mHV 30g). Inclusions mark the butt- and pattern-
welded rod - which was visible with the naked eye on the blade surface in the middle
of the blade body. The rod consisted of twisted mild and hard steel wires (ferrite-and-
martensite, martensite about 600 mHV). The cutting-edge, also butt-joined, is marten-
sitic, up to 1100 mHV; the cutting-line proper is softer (ca 600 mHV), possibly a result
of the heat developed during intensive grinding. The knife was a masterpiece, apparently
partially quenched along the cutting-line only (Pl. XXXVIII). Note: It is not uninter-
esting that a buckle-shaped strike-a-fire from the same site has shown a back behind the
striking edge which was twisted out of iron and steel wires as if it were pattern-welded
(Pleiner o.c. 273, pl. 50: 10 - 12).
The making of pattern-welded knives ceased sometime during the 14th century. Per-
haps it was too laborious in economic terms as the tendency was now towards large-scale
activities. Nonetheless, other styles which employed the sensible combination of iron and
hard steel through forge welding continued (Pleiner 1991a).
In modern time, the pattern-welding lives on either in experimental work or as a curios-
ity: certain specialist firms, eg. in Germany, produce different ware applying the principle
of pattern-welding; they edit their own catalogues.

Armour making

During the Middle Ages the production of armour, manufacturing protective means
against wounds, developed to a specific branch of ironworking. Highly skilled special-
ists delivered their artefacts for military purposes. Very abridged remarks are devoted to
this topic on following pages.

Notes on chain mail armour

One could hesitate to rank the chain or ring-mail armour amongst top technologies. With
some exceptions the rings were made of low carbon forged iron, rarely of steel wire (for
Roman specimens from Stuttgart see Fulford et al. 2004, 84, figs 1 and 2) but what
remains astonishing is the incredible fine work. The open ends of the rings of about 1cm
in diameter were alternately joined by welding and piercing and riveting (E.D. Schmid
2003). Kolchin (1953, 150 - 152, figs 122 - 123) investigated these elements from various
early Russian sites of the 10th century AD and estimated that a tunic-shaped ring-mail
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 223
short (kolchuga) consisted of 18 000 - 20 000 rings.
Traces of mail armour have been found at several the Late Iron Age sites. The Romans
wore loricae hamatae, ring-mail shirts and medieval armourers put together mail shirts or
parts of mail armour up to post-medieval times. The 10th century shirt of St. Wenceslas
desserves attention. It is kept in the cathedral treasury at Prague. The rings were made
of iron but the collar contains some chains of golden rings.
The entire Middle Ages witnessed the use of chain mail armour, also as partial sup-
plements like shoulder-pieces, collars and hoods. Even in the 17th century the Russian
governor of the former ‘Russian America’ (Alaska) wore a ring-mail shirt under his uniform
to be protected against any accident.
Joined chains of rings provided weaker protection against missiles, including arrows
and especially medieval crossbow-bolts. Because of that iron or steel bands and sheet
plates were gradually introduced to shield critical body parts.

Plate armour

Non-ferrous metal helmets, cuirasses and shin-guards are known since the Late Bronze
Age and survived until Classical Antiquity. The Roman army of the Imperial period was
armed with iron protection, helmets and cuirasses of iron scales (loricae squamatae) and
of iron plates or bands (loricae segmentatae). The latter are of special interest because, as
was the case with the bronze armour, the individual armour parts had to be adapted to
human body: skull, shoulders, trunk and limbs. To take anatomical details into account
in iron was an additional task for the armourers.
Recently, pieces of Roman armour from Rhineland and England were investigated
(Fulford et al. 2004, with references). This produced interesting information: apart rom
several steel mail rings, the sheet armour was iron (80% of 43 investigated specimens).
The iron sheets of the loricae were mostly heavily cold hammered to ¡1mm of thickness, the
ferrite grains were drastically deformed reaching 128 to 313 HV hardness. The technology
must have involved intensive work with sledge or set hammers; the authors consider even
a cold rolling of iron but cannot provide any evidence of a rolling device.
As regards the manufacture of medieval armour, the developments in the making of
helmets may be briefly outlined. Lighter conical helmets (often with nasal bars) were
supplemented with the knights’ so-called great helmets, the Topfhelm in German (the
13th - 14th centuries) which were mounted and riveted together from several metal plates
in which visors secured the sight. The weight was 2kg - 2.5kg. So far three helmets have
been investigated, the material was ferritic iron in two cases and ferritic-and-pearlitic steel
in one case (Williams 2004, 645 - 647, figs 1 to 6). About AD 1300 a lighter bascinet,
forged from one piece of sheet, began to be introduced. Bascinets without visor weighed
1kg - 1.6kg; however those with movable visors served as heavy head protection. The
example from Churburg, final 14th century, weighed 4kg (iron or very mild steel, see
Williams o. c., 648 - 649, figs 11 to 14).
The so-called sallet type of helmet was typical for the 15th century and was produced
in certain variants known from Europe. The lower part of the helmet was riveted to
the skull; the face was partly protected and sight was enabled by a visor. The rear part
defending the neck protruded as a tail. The technology has been recognized in an example
from Landshut, Austria (see below). The trunk of the warrior was protected by breast-
and backplates. In addition to the above mentioned helmet from Churburg a breastplate
224 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
has been examined (Williams o. c., 650, figs 15 and 16). The pearlite structure with
ferritic network shows that this piece of armour was made of steel containing up to 0.6
% C. It was forged from one piece, weighed 2.63 kg and the investigator supposes that it
comes from Italy (marked with a P, possibly Petraiuolo da Missaglia, ca AD 1385).
The investigation of plate armour is inseparably connected with the name of A. R.
Williams who initiated the establishment of this field of study. He has concentrated on
display armour which was produced in particular during the 15th century in Italy and Ger-
many. Florence and Milan housed workshops of celebrated armourers (e.g. Pompeo della
Cesa) and princes like Cosimo dei Medici ran their own armour smithies (Williams 1978,
1991). There were places in southern Germany (Nuremberg etc.) were places where fore-
most armourers also forged breast-pieces, helmets and other parts for the highest nobles;
the emperor Maximilian established, around AD 1480, his Court Armoury in Innsbruck,
Austria (idem, 1987, 1991b, with references). From the technological point of view, the
splendid armour was more frequently made of hard steel and heat-treated. The source was
bloomery metal or the newly introduced fined steel from the first charcoal blast furnaces
(idem, 1993, 2003). Some armours were ornated by gilding and other techniques which
upraised the appearance.

Example:

By courtesy od A. R. Williams there is possible to illustrate the production of two pieces


of armour which he investigated: a helmet bearing a municipal sign of Landshut, Austria
(Landshut B 129) and the armourer’s mark of Matthias Deutsch (ca AD 1490); then a
one-piece breastplate from Bayerisches Nat. Museum W 195 (ca AD 1400) cov-
ered with red velvet and bearing mark R (attributed to Jacomino Ravizza). The helmet
(sallet) is made of sheet steel (tempered martensite with a little ferrite), relatively free
of inclusions. The microhardness lies between 525 to 566 mHV 100g. The sample of the
breastplate shows a microstructure of very fine pearlite mixed with martensite and, again,
few slag inclusions. The breastplate may be termed as showy. Pl. XXXIX.
Note: It goes without saying that rank and file warriors in Europe were not dependent
on renowned armour workshops in Italy, Austria and Germany. Iron breastplates and
cheaper ‘iron hats’, kettle hats, were made in many urban workshops or even in smithies
in military camps and were accessible to large numbers of troops. For example, the war-
riors of the Bohemian Hussite movement in the 15th century were successful in many
battles and crashed the noble armour of the Crusader knights with their spiked flails and
iron balls which were fastened with chains to wooden helves (in Czech řemdih), even with
balls shot from the simple cannons of the time. Thus, these battles were of an unconven-
tional type and did not respect the rules of chivalry. Top class armour technology failed
in conflicts with national, social and religious ideology of the Hussites.

Locksmithing

The manufacture of locks was a specialized craft which served to safety close rooms and
containers. To a great extent it applied a precise coldwork in iron: grinding and filing.
In antiquity and in the Middle Ages Two basic systems have to distinguised in antiquity
and and in the Middle Ages: padlocks and fixed revolving locks.
Padlocks were already used in Roman times (some examples have been mentioned
METALLOGRAPHY OF EARLY IRON: RECONSTRUCTED TECHNOLOGIES 225
from Britain). They were frequently used during the early Middle Ages in urban cul-
tures. Kolchin (1953) has paid considerable attention to early Russian padlocks (more
than 300 have been uncovered in archaeological layers of Russian towns of the 10th -
13th centuries). The cylindrical bodies were made of iron but the functional parts, the
bow-shaped inserting pegs, demonstrated sophisticated technologies what concerns the
fastening of two oposite springs to the central stick by soldering with soft or semi-hard
solders (Pb, Sn, Cu). The springs of individual locks were made of iron or tempered mild
steel, the opposite ones sometimes of tempered hard steel up to ca 0.75 % C (ibidem, 161,
based on 12 examinations). Special keys with a perforated terminal disc were needed for
unlocking. Padlocks were produced in town locksmiths’ workshops but occasionaly got
into the countryside where they could have been repaired by local smiths (Mutějovice,
Bohemia, 13th century, see Pleiner 1969b, 556, 560, fig.16, smithy A).
No less precise work represent the early fixed revolving locks of doors or lids of chests
and strong-boxes were equally precise workmanship. Toothed keys turned the bolts of
lock forwards and backwards and closed and opened the mechanism like in modern times
(Rasl 1988; 1995). Unfortunately, no analyses have been undertaken on medieval fixed
locks.

Clock making

Medieval clock making was another top class craft. It involved the solving of time mea-
suring problems. From the mechanical point of view it required, as in the case of lock-
smithing, laborious and precise abrasive coldworking processes. During the 14th century
tower wheel clocks, medieval horloges, functioned using the motive power of gravity on
weight and balance beams. The 15th century horloges in many towns were mechanical
masterpieces. No metallographical tests of horloge part have as yet beens carried out now.
Chapter XII

CONCISE HISTORY OF EARLY EUROPEAN BLACKSMITHING

The problems connected with the beginning of smelting iron ores have been discussed
in the author’s previous book on bloomery smelters and there is not intended to repeat
the facts, theories and hypotheses on these pages. Instead, attention shall be concentrated
on practical and symbolic use of the black metal in time and space up to the Middle Ages.

The beginnings of the use of iron and the first smiths

In the earliest times, the ancient metalworkers could obtain limited amounts of the new
metal, precious iron, from accidentally produced pieces in the copper-smelting processes
applying chalcosiderites or iron ore fluxes or from intentionally smelted iron ores. The
third source were iron meteorites. The frequency of iron of cosmic origin is unclear
because the older finds, declared as of meteoritic origin due their elevated nickel (and
cobalt) contents have recently been regarded as dubious, since nickel concentrations, in
the welds, for example, can be high, when the iron was smelted from nickel-bearing iron
ores. However, there are finds, which have recently been examined by modern chemical
and mineralogical methods, which are definitely of meteoric origin. So, the third source
can certainly not be elimited.
At any rate, some of the metalworkers, usually in the service of prominent members of
society, got hold of some iron and were able to work it. Apart from scarcely identifiable
objects like various beads and corroded lumps, the finds show some rings and pins dating
from the 3rd - 2nd millennia BC but, what is important, first daggers with iron blades
appeared (Tell Asmar, Alaca Höyük in Anatolia, Ur in Mesopotamia, Thebes in Egypt).
At the end of the 3rd millennium iron (kù.an) was mentioned in legal texts and epics
in the Near East and precious iron daggers and axes are said to have played a role in
ceremonies like oath swearing (Shusin IV, Lugalbanda epic).
The uncontrolled process of the early smelting of iron ore mostly produced a hetero-
geneously carburized metal. However, sometimes hard steel was worked, as the (blade?)
from Pella, Jordan shows; this dates from the Middle Bronze period. Apparently, it was
an exception; in the early days of ironworking carbon-poor or heterogeneously carburized
iron was the material for smiths.
At that time, the amounts of iron available and used for making artefacts could be
measured in grammes; this could be characterized as a shekel iron economy (a shekel
comprised about 16 grammes and a ca 20cm long dagger weighed some 5 shekels or 80g).
Later, the Near Eastern Bronze Age witnessed the appearance of swords, maceheads,
axeheads an the like the weight of which could be calculated in minnae, about 0.5kg
each). In the temple of Guzana-Tell Halaf in Mesopotamia a curious movable firegrate
was discovered, made of 1m long iron rods fixed on an undercarriage with small bronze
wheels; it dates from about 1200 BC.
In terms of the use of iron, Europe lagged behind. The first samples of iron came to
light at about the same time but merely as sporadically occuring objects (minute rods,
rings). The example of an iron dagger hilt from a ritual well at Gánovce, N Slovakia, was
apparently a 15th/14th century BC import from the Greek Mediterranean. Several other
CONCISE HISTORY OF EARLY EUROPEAN BLACKSMITHING 227
isolated small iron finds dating from the Hallstatt A period are known. Central Europe
witnessed an increasing frequency of the black metal during the final Bronze Age (9th/8th
century BC). More than 50 items as knives, swords, lanceheads, axes and ornaments and
even blooms (2 cases) appeared among the finds but iron remained a prestige commodity.
In contrast, the Near East entered the fully-fledged Iron Age at that time. In the 8th
century BC, iron circulated (especially as spoils and tributes) in tonnes or thousands of
talents (ca 26 - 30 kg each) in regions influenced by Neo-Assyrian rulers as attested not
only by inscriptions but also by the famous store of iron at Khorsabad-Dur Sharrukin. In
ancient Greece, the developed Iron Age started as well.
The iron objects represent the work of craftsmen who were starting to master iron-
working, metalworkers from whom iron smiths gradually separated. They are mentioned,
for the first time, at courts of Hittite kings in Anatolia. This cultural and political sphere
of the 17th to the 13th centuries BC can, in fact, be considered as the cradle of iron mak-
ing, preceding the formation of the Indo-European Hittite confederation and introduced
during the era of the non-Indoeuropean Hatti.
Although the relevant archaeological finds in Anatolia remain meagre, the work of
blacksmiths is reflected in numerous cuneiform documents from the royal archives at
Boghazköy-Hattusas. Ritual texts mention symbolic and even surreal notions (heaven of
iron - AN.BAR, mountain, tongue, words) alluding to power and strength. Then, iron
as a component of cultic objects: vessels and their lids, human and animal statuettes,
ceremonial lances, clubs etc. Other documents, those of the profane or secular category
(contracts, edicts, royal letters etc., especially from the later Hittite period) point to
weapons (daggers as royal gifts in the famous Hatussil letter KBo 14) and various blades
and even ingots. The creators of these artefacts were specialists: smiths. Certain texts not
only speak about iron smiths (LÚM ES AN.BAR.DǏM. DǏM) but also about their chiefs
and overseers. The smiths and blacksmiths were evidently dependent people serving
at local courts, delivering their products and taking the subsistence. Meteoritic iron
(AN.BAR.GE6 ) was occasionaly worked to clubs and weapon blades.
After the fall of the Hittite realm (about 1200 BC) the strict confinment of these
craftsmen was loosed to certain extent. The possibilities of migration come into question.
However, they were specialists keeping the secrets of their art within their families. They
were still not able to apply it extensively but only on behalf of the prominent groups in
society. Because of that they again became, ‘court craftsmen’, dependent workers.
What technologies did the earliest smiths master? All of those necessary for hot (and
cold) forming up to sharpening, pointing and planishing. As concerns the influence on
the properties, it depended on the accidental or deliberate choice of the material and
attempts to meliorate it, the information is supplied the metallography; the number of
examinations is still relatively limited. Exceptionally hard steel has been observed (the
object called blade from Pella, Jordan, 2nd millennium BC, which was quenched, be
it accidentally or intentionally; the all-steel knife from Radzovce in Europe, Slovakia,
8th century BC). The majority of the objects were made of softer iron, some of them
were heterogeneously but slightly carburized. Among the weapons and tools from the
Near East dating from around 1000 BC pieces were interpreted as secondarily carburized
(ploughshares, blades, punches) but agressive corrosion has hindered the presentation of
a clear pattern of structural arrangement (Taanach in Palestine and other localities in the
eastern Mediterranean). Heat treatment (quenching, tempering) has been reported from
Pella and from Idalion, Amathus and other sites in Cyprus, in the case of knife blades.
228 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
As for Europe, the sword blade from Singen, SW Germany, 8th century BC, demon-
strated the welding-together of two main iron bands; a hard steel plate emerged near the
hilt part but no system could be identified.
The smiths of these early periods used the kind of metal available and began to ex-
periment with melioration techniques (e.g. additional carburization). The technological
processes occurred in embryonic state.

Blacksmiths in the ancient civilizations of Greece and Rome


The classical civilizations of Greece and Rome started the spread of advanced ironworking
and influenced this field in the rest of Europe.
Leaving aside the Greek mythological tradition concerning ironmaking and -working
groups (Daktyloi, Korybantes, Kuretes, and more real Chalybes), the ironworking prac-
tices were introduced via Cyprus the mediating role of which has to be duly appraised.
Ironworking was represented, as archaeology has shown, by cutlery, the fabrication of
knives during the 12th to 9th centuries BC; in the later period of this epoch the Aegean
was the scene of a similar development. The symbolic significance of preceding finds
of Peloponnese (parts of iron rings and amulets) does not indicate any changes in this
process.
The Homeric complex of Iliad and Odyssey contains about fifty places where iron
(sidéros) occurs. The compilation which was finished in the 8th century BC involves
several metamorphoses rendering the metal as a sign of strength and power, and certain
material forgings are commented on: a mace and arrow, an axe. In general, iron is
presented as a precious metal, a welcome prize in races and a commodity of wealth.
Within the final version of the poems a stratification in time of these places has to be
respected: some of them are related to the siege and conquest of Troy before 1000 BC
but other must have been contamined by later development. This concerns the passage
on the heavy iron discus of Achilles who described it as a source for making ploughs.
This is distinctly marked by the experience of the full Iron Age. The same may be said
about the important place in Odyssey where, in the tale of the blinding of Polyphemus,
the quenching of a heavy iron axe is described (Od. IX 391); however, some of the above
mentioned Cypriot and Aegean knives were quenched or tempered as has been attested
by Swedish metallographers.
The development during the late Protogeometric and Geometric periods (10th to 8th
centuries BC) is marked by the gradually increasing role of iron pointing in the fully-
fledged Iron Age. Apart from iron ornaments (fibulae) iron weapons like swords, lances
and axes were common. A wide range of craftsmens’ tools and farming implements came
to use (incl. iron saws as well). Iron sheet could be hammered as toreutic ware and applied
in its production. Bulky objects were forged of iron (massive oil lamp stands, fire-dogs).
More than thirty kinds of forgings are attested both by archaeology and written sources.
The latter involves the evidence yielded by Hesiod (7th century BC) - in the Theogony
he was the first to call his own era the Iron Age (in his view the fifth epoch in the history
of mankind). He antedated the chronological system of Christian Thomsen (1788 - 1866)
by more than twenty five centuries. More material information is contained in his ‘Works
and Days’; iron artefacts fell trees and served in any kind of woodworking, ploughing,
harvesting and, of course, fighting. The workshops of master smiths became refugees for
people coming to hold talks.
CONCISE HISTORY OF EARLY EUROPEAN BLACKSMITHING 229
About the same time another phenomenon affected the culture and economy of ancient
Greece: the smiths produced stylized implements used as means of exchange and even as
currency: they are named in written records and known as archaeological objects as well:
anchors ankýrai, sickles drépana and above all spits obeloi, obeliskoi which gave the name
to the money. This system lasted several centuries up to the 4th century BC preceding
struck silver coins. However, struck iron coins (sidareioi) circulated in several city states
of that century. Hard striking tools must have been used in their manufacture.
During the Classical and Hellenistic periods more than fifty kinds of iron artefacts
served the inhabitants. From the 5th century BC it was possible to apply iron production
in the form of structural clamps joining the ashlars. Metallographic examinations have
shown that two T-shaped bars were welded-together into a H-shape and sealed with lead
in special cavities in the building of the renewed Parthenon temple in Athens.
Not only obeloi but also other iron objcts came out of circualtion as votive gifts from
groups of population after their consecration in temples.
Naturally, considerable differentiation evolved in the technical ability of the black-
smiths. Xenophon remarks that rural smiths were less experienced in making implements
that the masters in the town. There, in large smithies worked specialized craftsmen,
swordsmiths: machairopoioi, xifourgoi, cutlers (drepanourgoi), hoe-makers (sminyaspoioi).
They were high value slaves. Demosthenes (around 300 BC) recalled ”thirty two or thirty
three swordsmiths in a sword workshop owned by his father, a man of the jury”. Factories
producing shields employed numerous craftsmen (up to 120 in Piraeus) and cuirasses of
the best quality were reported (Zoilos, a skilled armourer, worked about 300 BC). For the
most part, the information on ancient Greek ironworking comes from written documents.
Artefacts uncovered by archaeology play a rather complementary role. As to the Hel-
lenistic period, the finds of the Greek Priene on the western coast of Asia Minor deserve
a note, as well as iron picks and sledges from mines at Thasos and Laurion.
The technological processes mastered by blacksmiths have not been properly elucidated,
since no systematic metallographic examination has yet taken place, with the exception of
some spearheads and clamps; more information could be obtained with the aid of about
30 samples (axeheads, adzes, hoes, swords) from the Greek Black Sea colonies, especially
from Gorgippia (mostly the 3rd century BC) on the Kerch Strait. Secondary carburization
and welding-together of iron and hard steel as well as heat treatment (quench-hardening)
have been observed indicating the high level of the technology. Aeschylus and Sophocles
mention the quenching of steel (bafé sidéros) and Hippocrates knew about tempering in
oil (sidéros bafénto eis elaion).


Several iron objects (rings, blades) appeared in Sicily and southern Italy (Catania) in
about the 10th century BC. The origin of these items remains murky, although Aegean
influence certainly comes into question. More finds date from the 8th century BC and for
that period the Greek coastal colonization also has to be taken into account. The Greek
poleis on the western coast contributed to the process. In Pithecoussai on Ischia island
smithies from the late 8th century BC have been uncovered.
The spread of the use of iron among Italic tribes into the inland of the Apenine Penin-
sula has to be connected with the activities of Etruscans (Tyrrhenoi in Greek, Tusci,
Etrusci in Latin, Rasenna in their own language). The origin of this ethnos is under
230 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
discussion. The most logical opinion is that of the ancient Greek historians who held
this people to be a branch of Anatolian Lydians who had migrated to Italy by sea. This
would explain their knowledge of ironmaking (large-scale smelting of Elban iron ores on
that island and at the adjacent coast of Tuscany) and iron forging. However, Dionisios of
Halicarnassus thought at his time (as late as around BC/AD) they were a local people.
(The theory of the coming of the Etruscans by land across the Alps seems to be highly
inadequate). As a matter of fact, the Etruscans in Italy had been exceptionally familiar
with the working of iron since the 10th century BC. By the 8th century BC they were
forging swords and spearheads and were spreading the knowledge of the use of iron from
their territories to all directions into the inland.
Generally, the Iron Age in Italy was established during the 7th to 5th centuries BC.
Subsequently, progress must have been rapid, although there is lack of direct evidence.
A considerable increase in the iron industry during the later period of the Roman Re-
public benefited from the Etruscan production of iron on Elba and in Tuscany which
was apparently under Roman control. An early evidence of blacksmithing is supplied by
Pompeii having been destroyed by the eruption of Vesuvius volcano in AD 79. Several
smithies were discovered during old and resent excavations; as a rule they were located
at gates. The smiths worked iron and bronze as well. A shop was uncovered where tools
and implements were tendered.
Since the 2nd century BC Italic entrepreneurs started ironmaking in southern France
(Gallia Narbonnensis) and in the Roman provinces where the evidence of blacksmithing
can be traced, be it in very early legionary castra in Spain (Hispania) or be it in Roman lay-
ers in numerous provincial towns, villages and estates. The smithing tools ranked within
the developed category: hand hammers (often of a ‘Roman style’ with edged sides at the
circular shaft holes), sledge and set hammers, hinged tongs (universal and with shaped
jaws), piercing and riveting irons, block anvils with nail holes in a corner or horned anvils
adapted to bending operations, swages and dies, wire drawing irons, metal sheet shears,
forge spoons. Smithing tools were the equipment of smithies but they rarely outlasted in
their remains. Archaeology has a number of excavated workshops at its disposal, not only
in towns where many of them have been explored in Gallia Narbonensis, in Britain, in
Serbia (Maslovare, specialized in nailmaking). In Augustodunum-Autun, France, an arti-
sanal district housed several stone-walled workshops. Smiths worked in villages (Marsens
in Switzerland) and in the villae rusticae (Biberist-Spitalhof in Switzerland, Kervenennec
in Brittany, Petöháza in Pannonia, etc.). The plan shape was different and covered the
room of some 16m - 40m. Other smithies functioned at roads (e.g. Noreia or a number of
excavated workshops in the southern environment of Alesia in Burgundy) rendering ser-
vices to travellers. As yet, no archaeological evidence has become available concerning the
large Roman weapon and armour factories known from the Notitia dignitatum of the late
Imperial period. Those from Augustodunum-Autun which produced armour, arrows and
siege machines are not identifiable amongst the civil workshops in the artisans’ quarter
mentioned above. Such fabricae existed in France, Italy and Roman Balkan provinces. In
central Gaul, a ‘fabrica armorum omnium’ at Argentomagus near Bourges must have been
an admirable manufacturing outlet. Recently, metallographic examinations on about 40
Roman armour parts from England and north-west Europe stand at disposal (helmets,
shield bosses, body armour of ring mail - lorica hamata, scale armour - lorica squamata,
banded armour - lorica segmentata). Steel was used in a minority of cases (one ring of
mail was even quenched), mostly iron was applied which was heavily cold-hammered, very
CONCISE HISTORY OF EARLY EUROPEAN BLACKSMITHING 231
hard and poor in inclusions. The lorica bands were very thin, less than 1.2mm thick. The
provenance of the investigated armour parts is not known.
The smiths used various stock as starting material, scrap, bars, rods and ingots. The
latter involved either small quantities of pure iron (stamped) or a hard steel, also stamped
by the name of the smith (e.g. Saturninus). Iron and steel were traded (Norician steel,
sunken ship cargoes of this commodity in the shape of bars in wrecks in the coastal waters
of Southern France and elsewhere). The masters of ironwork were able to supply peo-
ple with all kinds of implements for crafts and material tretament, cultivating arable land
and household equipment and domestic utensils incl. padlocks, for example. Swordsmiths
could offer short swords (gladii) and officers’ daggers (siccae) of good quality and the ar-
mourers cuirasses (loricae hamatae, squamatae, segmentatae). Irons applied in structural
works deserve particular note: large nails (about 5 tonnes were buried and sealed in a
pit at the fort of Inchtuthil in Scotland so as not to get into enemy hands after the site
had been evacuated). As in Persia and Greece the ashlars in monumental stone buildings
were joined with iron clamps (e.g. Rome, Porta Nigra in Augusta Treverorum-Trier which
later fulfilled the role of a mine yielding ready-made iron). The builders of Roman baths
used heavy forged beams supporting the heating installations; they weighed up to around
120kg. These were found in several sites in Britain and western Germany.
Except for the objects abandoned in ruins a large amount of forged iron hiding valuable
items. Apart from Inchtuthil finds of such a kind were uncovered in the Roman military
stations along the limes. The reason for depositing large hoards of iron artefacts could be
similar (Selz, Kreimbach, Newstead, Dombóvár and others). A votive intent seems rather
unlikely.
All this production led to the specialization of blacksmiths. Some of the civil branches
are made up of toolmakers armamentarii, cutlers (cultrarii, falcarii), hoemakers (dolabrarii),
locksmiths (clavicularii). The social position of blacksmiths was extremly varied: slaves,
franchised libertini, citizens (many of them were wealthy and could leave well-built tombs
with symbols of their craft); and, of course, military smiths in army posts on the frontier
and inland.
The standard of Roman technology, as regards metallography, was heterogeneous. Re-
sults achieved from the relatively few analyses on artefacts from Germania Superior and
Agri Decumates (on the left and right banks of the middle Rhine), and from Rhaetia
(Bavaria), show a lot of carbon-poor implements and those made by more progressive
techniques with application of hard steel. The smiths from the latter site achieved per-
fection and even top quality in the manufacture of weapon blades many of which were
pattern-welded and some examples crossed the border as spoil or gifts to Barbarian lands.
The characteristics of Roman blacksmithing were the wide range of qualities as regards
to the ware and, above all, in the contemporary European world, the quantity.

Celtic ironworkers
The development of Celts inhabiting south central and western Europe took place syn-
chronously with that of Rome and Celtic tribes were in intensive contacts: hostile as
well as peaceful (Celtic raids even endangering the Capital). The Romans were forced
out of northern Italy - Gallia Cisalpina by the settlements Celtic tribes but the region
was romanized again during the 3rd to the 1st century BC. This resulted in some sim-
iliarities in lifestyle - the occasional exchange of goods and, above all, of the know-how.
232 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Pliny (Hist. Nat. 12.2.5) has recorded a tale in which a Celtic blacksmith named Helico
ex Helvetiis settled in Rome having been invited by his countrymen. The Celtic smiths
produced swords, lances, shield fittings, not to mention knives, personal ornaments and
the like just in the period around 400 BC. No metallographic investigations have been
undertaken so far. On the other hand, more than 130 Middle La Tène Celtic swords have
been examined. No standard techniques were recorded: the smiths produced all-iron and
all-steel blades, piled examples, or used secondary carburization of cutting-edges or joined
hard steel to the blade core (shells) or welded in a steel band in the middle (sandwich).
There are rare cases of pattern-welding (Cleebronn, Cuvio). No quenching was observed.
Making of metal scabbards, often richly decorated, was part of the swordsmith’s work.
The Celtic panoply involved massive lanceheads and shaft butts, chain belts, shields fit-
ted with iron rims and bosses. Ring-mail armour occasionally came light not to mention
horse-gear and fibulae.
Dynamic progress followed during the 2nd and 1st centuries BC. This was partly in-
fluenced by intensive contacts with Classical world, partly by the inventive nature of the
Celts. About 90 kinds of iron forgings were left in settlement layers, mostly in centres of
the oppidum type: bars, ingots, tools, ornaments, domestic utensils, weapons and parts
of armour. Large and heavy iron nails have joint the wooden beams of the murus gallicus
type of the fortification. Tonnes of iron were required for this. Hand hammers (some of
them of steel and quenched), block and horned anvils, chisels, files, forge-spoons occured
among the smiths’ tools. The smithies were light post-huts or shelters. Traces or ground
plans are mostly known from the oppida (Bibracte-Mount Beuvray, Rheinau, Manching,
Závist, Hrazany and Lhotice in Bohemia - the latter having produced iron fibulae). The
sunken-floored smithy at Mšec, Bohemia, was part of a smelting complex. A fully-fledged
iron civilzation emerged.
The neighbourhood of advanced cultures, radiating the knowledge, and the own skill
of Celtic masters, were a fertile field for the beginning lounch of ironworking techniques.
Five artefacts (4 axeheads) have been metallographicaly examined from 7th to 3rd century
BC Ireland; surprisingly their blades were welded from iron and steel. In spite of the
wide date range they could be of early Celtic origin. There are about 80 metallographic
analyses from the central Celtic territories north of the Alps. They mostly come from the
oppida in Bohemia, Moravia and Slovakia, and from Manching in Bavaria. About 30% of
tools were entirely made of a relatively soft iron, about 20% represent all-steel forgings,
about 18% were secondarily carburized and more than 30% were welded-together of iron
and hardenable steel (in sum, one third was heat-treated by quenching or tempering).
Progressive techniques were also discovered in Iron Age Britain (no statistics can be
presented as only some results have been fully published whereas others have only been
issued in a preliminary fashion). A different situation reigned on the northern periphery
of the Celtic domain, e. g. in Thuringia and Silesia where several dozen investigated
forgings have been published or reported. Only a low percentage of the samples revealed
the intentional iron-and-steel welding, all-steel blades respresent about 10%; the bulk was
manufactured of iron or unhardenable steel (66%).
The material (stock) for making iron things involved, apart from scrap, bipyramidal
ingots on the one hand, and bars (taleae ferreae) resembling stylized tools on the other.
The first type has been principally preserved in hoards. About one hundred have survived
comprising around 700 pieces. The deposits are concentrated in Bavaria, Württemberg
and Hesse, the homeland of the Celtic ethnogeny. The bars were each welded-together
CONCISE HISTORY OF EARLY EUROPEAN BLACKSMITHING 233
from several iron blooms. This form can be traced from the late Hallstatt period up to
2nd/1st century AD. The second kind of stock entered the early history as the ‘currency
bars’ (Caesar, BG 12.5), especially in the Iron Age Britain. They come from hoards as
well as from settlement layers. The reason for burying hoards might have been different
but the enormous quantity of the bipyramidal ingots gives rise to the suspition that we
are dealing with the ideological sphere - votive gifts. It seems that the religious-minded
Celtic folk simply took risks with valuable material and that the losses of iron, taken
out from the everyday use, might have caused or contributed to the serious defficiency
and, finally, collapse of the Celtic economy despite the fact that Celtic blacksmithing
ranks amongst one of the most progressive traditions over a great part of Europe. This
tradition continued in Gaul where the Celts in the Gallo-Roman epoch assimilated with
the colonists. Their crafts prospered.
Far in the west, in Ireland, the Celtic culture of the Early iron Age B (200 BC to AD
500) naturally inwolved iron working. Around thirty objects were investigated. Simple
techniques were observed in the majority of cases.


In south-eastern Europe, in the territory of modern Romania, eastern Hungary and south-
eastern Slovakia, the neigbours of the Celts were Dacian and Getian tribes, both of Thra-
cian origin, who mastered the art of ironworking as well. Unfortunately, no synthetis
dealing with this theme exists. However, the huge hoards of iron bars in the capital
centre of Sarmizegetusa (modern Dealul Gradiştei or Gradiştea Muncelului indicate that
iron metal like silver and gold, was an important material which lured conquerers to these
countries. After AD 106 Dacia became a Roman province.

Beyond the Roman frontier: the Germanic, Cimmerian and Scythian tribes
Since the prehistoric times, the north of Central and Scandinavian Europe was the scene
of the development of Germanic tribes. In the areas immediately to the north of the
Celts Germanic tribes and tribal complexes entered the awareness of classical writers as
e.g. Hermunduri, Thuringi, Marcomani, Quadi, Lugii, Suebi and many others, having
inhabited Scandinavia and what is modern Germany and a part of Poland. Around the
dawn of the Christian era they entered the Celtic lands and became immediate neighbours
of the Roman empire. Iron had been worked by these peoples since later Hallstatt times
into basic categories of artefacts, basic weapons and tools which were generally manufac-
tured by simple techniques from available materials:unevenly carburized mild steels. The
Germanic peoples encountered Rome with this equipment.
The Romans tried to conquer the western tract up to the Elbe river (Cherusci, Chatti,
Longobardi etc.) but their efforts failed in many military campaignes. They established
provinces called Germania Inferior and Germania Superior which only absorbed a cou-
ple of Germanic (and Celtic) groups on the left bank of the Rhine river. The Rhine
(and Danube) began to be watched and the unclear border between two civilizations as-
sumed distinct outlines and remained in existence practically four centuries. It began
to be fortified from the late 1st century BC. Lines of military roads connected castella,
castra, burgi and crossed the triangle between the upper streams of Rhine and Danube
(Agri Decumates). The Roman limes played the role of an ‘iron curtain’ dividing the two
234 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
worlds which occasionaly breached by raids of Germanic groups. Some goods also came
over the border (spoils, gifts, including e.g. iron swords) and limited trade formed the
starting point for commercional routes like Regina Castra-Regensburg and Carnuntum
east of Vindobona-Vienna. What did not cross the border was the artisanal know-how.
This can be illustrated just on hand of ironworking technologies. As mentioned, the Ger-
manic smiths continued to work carbon-poor iron and heteogeneously carburized steel into
weapons or tools. The great deal of metallographic analyses which have been undertaken
especially in Poland demonstrate that the application of more advanced techniques was
rare. This has been shown verly clearly by examinations of artefacts buried in graves of
several Germanic cemeteries in southern Slovakia which are located just within eyesight
of the Danube Limes.
The only sign of a penetration of Roman elements of industrial activity comes from
the Holy Cross Mountains area in central Poland. The large-scale smelting of an excellent
iron ore started in the La Tène period, increased in scale around the 2nd century AD
and possibly attracted may be individuals, may be groups of interested people connected
to Rome (underground mining, Roman coins). Several iron tested iron objects (knife,
lancehead, punch etc.) have shown iron-to-steel welding and simple techniques as well.
The separation of the two worlds, the Roman and the Barbarian, depended not only
on account of political reasons but also by the difficulties of land traffic. This changed
due to maritime radiating of Roman technical progress via Britain to Scandinavia and
Baltic. During the Vendel and early Viking periods and later in the Middle Ages perfect
technologies in iron began to flourish in these parts of Europe. The picture of the devel-
opment of ironworking assumed quite different outlines.

The wide plains of steppes and steppe-forests north of Caucasus and the Black Sea possibly
adopted iron during the 2nd millennium BC from two directions: from eastern Anatolia
along the eastern coast of the Black Sea, and from western Anatolia along the western
coast of the ‘Pontus Euxinus’ towards the North-East.
The first users of the still rare iron were the bearers of the Koban culture north of
the Caucasus (the flourishing phase of which covers the 11th to the 8th century BC).
Their predecessors, presumably the early Cimmerians, an Iranian people, were acquainted
with meteoritic iron worked into bi-metallic tools (Boldyrevo, 18th century BC). Later,
Cimmerian blacksmiths produced, for example, nice daggers with cruciform dagger hilts,
awls and trunniom axeheads; the movements of these tribes started around 800 BC. They
were caused by the pressure of the coming of the Scythians, another Indo-European people.
Cimmerians invaded Asia Minor on one side, and the Lower Danube regions on the other
and iron forgings were a part of their equipment. The role of the so-called ‘Cimmerian
iron route’ to central Europe does not seem to have been important merely bringing some
of the above mentioned cruciform daggers. More than 150 metallographgic analyses show
all-iron, all-steel (often superficially decarburized) artefacts; occasionally edge-carburized
specimens (trunnion axes); the welding-on of steel was observed on knives, sickles and on
a sword.
The greater part of the territory of the steppe and steppe-forest regions of modern
Ukraine and southern Russia were occupied by the Timber Grave culture (the later phase
CONCISE HISTORY OF EARLY EUROPEAN BLACKSMITHING 235
of the Ochre Graves) which had developed since 1500/1400 BC. The iron inventory in-
cluded awls, knives, and from the later centuries bi-metallic daggers as well (11th/9th
centuries BC). After that, the archaeological finds are connected with Scythians, who
expelled the Cimmerian tribes having been endangered by other nomads, the Sarmatians
(Sauromates by Herodotus) from the east. Scythians were nomads as well (Royal Scythi-
ans) who developed a culture involving perfect bronze founding, silver- and ironworking
as attested by finds in huge burial mounds, presumably also having been influenced by
Greek colonies on the northern shores of the Pontus Euxinus (Olbia, Tyras, Pantikap-
paion, Phanagoria, Gorgippia, the latter yielding examples of an advanced ironworking
techniques). However, important centres appeared in this nomadic milieu like the Belskoye
complex (Gelonos) and Kamenskoye where, among other things, metalworking flourished.
Herodotus was also aware of southern Scythian tribes engaged in agriculture (Aroteres -
tillers, and Georgoi - farmers).
Scythian smiths forged at least twenty kinds of artefact, above all weapons, swords,
lances and arrowheads. The latter were cast in bronze but iron examples are also present.
Their finest achievement was the creation of a short sword or dirk called akinakes with
a heart-shaped guard. An early example (said to be from the 6th century BC) was of
mild steel but both edges were carburized up to 0.7% C. Other pieces were made as all-
steel blades or submitted to secondary cementation. The use of heat treatment has been
attested on knives and dirks. These akinakes in splendid golden or gilded scabbards richly
ornated with animal motifs come from imposing royal burial mounds (kurgan). Their
producers must have mastered the jeweller’s work or cooperated closely with jewelers.
The Scythians drove carts and four-wheeled vehicles. Fragments of nailed iron tyres
have been found and examined. All in all Scythian ironworking has to be classified as a
developed craft.
The Sarmatians - Sauromates, settled the regions north of Caucasus and east of the
Dnieper river after the Cimmerians have moved. Their iron equipment was similar. For
example, they also wore the akinakes swords or dirks. They have absorbed the techno-
logical knowledge from the same background as their neigbours.
Subsequent history in the 1st millennium AD is marked by further migrations of people
and by a varied ethnic picture and reflects the different application of ironworking tech-
niques. The lands between the Dniester and Dnieper rivers were settled during the first
half of that time by tribes whose material manifestation were the Zarubincy, Kiyev an
Tchernyakhov cultures: Dacians, Sarmatians, early Slavs and Germanic newcomers from
the Vistula estuary - the Ostrogoths of the 4th century. Ironworking as seen on hand of
more than 220 investigated samples belonging to the Zarubincy culture shows nearly two
thirds of all-iron forgings, less than 20% all-steel with the welding-on of steel only occuring
exceptionally. The Kiyv culture yielded half or all-iron items but a complete third was
made up of all-steel objects and one tenth of blades with welded-on steel cutting-edges.
The Tchernyakhov culture: less than 27% all-iron, 33% all-steel and over 11% welding-on
steel. About one third of hardenable steels were quenched and tempered in all of the
named cultural complexes. It is very difficult to look for traces of an internal develop-
ment within the individual cultures or for streamings and influences in the technological
know-how.
236 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
European blacksmiths in the Early Middle Ages

After the fall of the Roman Empire the production of iron in Europe temporarily but
substantially decreased. The level of ironworking did not suffer as seen on the weapons
from Germanic graves of the 5th - 6th centuries AD. The spatha type long swords, saxes
and scramasaxes were of a good quality and were often plated with pattern-welded strips.
The self-etching caused dark and light tints of steel and iron components. Cassiodorus, the
personal secretary of Theodorich the Great who ruled in Italy AD 493 - 526, left in Variae
V.1 a description of a pattern-welded sword in a letter to a king of the Varni, a proto-
Slavic tribe in the north: ... harum (spatharum) medie pulchris alveis excavata quibusdam
videntur crispari posse vermiculis: ubi tanta varietas umbra conludit ut intextum magis
credas variis coloribus lucidum metallum (paraphrased: in the case of these [spathae] it is
possible to see nice central depressions with fancy crispations. There is such a variety of
iridescent shadows that you believe that various colours have been interwoven in the light
metal). Pattern-welded bands not only emerged in sword cores but also on splendidly
ornated Scandinavian and Baltic lanceheads. About AD 800 excellent swords of the
Ulfberht school appeared in the Rhineland and were apparently copied up to Sandinavia:
steel blades wore the ULFBERHT inscription as inlays in pattern-welded rods. Other
names of swordsmiths are known from that time (INGELI, +LEUTERIL+ in England,
+NISOMEFET+ and BENNOMEFECI in Germany, 11th - 12th centuries). A huge
effort of workmanship was required in the production of the mail armour (brunia). A
shirt consisted of about 20 000 riveted and welded rings made of forged wire. No wonder
that it must have ranked amongst luxury goods. Weapons and armour were sold for huge
amount of property: a sword and scabbard for seven cows and a sword and horse had the
value of a knight’s estate.
The lack of metallographic analyses on contemporary everyday-ironwork prevents us
from evaluating the general level of the blacksmith’s work. This is possible to do by means
of sources from later centuries (the 6th to 11th/12th) in Scandinavia and the Baltic and
adherent regions. Sophisticated iron-and-steel welding and effective tools construction
(incl. the three-layer sandwich system) and heat treatment were applied. The Anglo-
Saxons and the Irish mastered blacksmithing as well. Certain researchers have proposed
that fined cast iron which was obtained under specific reduction conditions in bloomery
furnaces was worked (Hamwic in Southern England).
A completely different situation and an admirable subsequent development can be
recorded among the early Slavs (Penkivka and Prague type cultures in the Ukraine - the
6th century AD). The ware of their blacksmiths was modest as over one hundred analyses
from the eastern regions have demonstrated. The situation was similar in central Europe
which was colonized by Slavs at that time. Just in the rural site of Dessau-Mosikau in
Germany the later phases have shown some well worked knives. At that time raiding
nomadic Avars had sabres at their disposal that have been with effectively carburized
along the cutting-edge or even constructed in a sandwich-scheme; a pattern-welded knife
appeared (several specimens of iron forgings were analysed, dating from the 7th - 8th
centuries AD; heat treatment was mastered). The source of the knowledge is difficult to
guess. The influence of the Khazars (Saltovo-Mayatsk culture in the Ukraine, 7th to 10th
century) comes into question. The Khazar kaganat subsequently became a part of ancient
Russia. The smiths mastered all the simple and devoleped techniques also forged sabres;
more that one hundred metallographic analyses have been published.
CONCISE HISTORY OF EARLY EUROPEAN BLACKSMITHING 237
Western Slavic culture moved a step forward during several generations so that Megalé
Moravia - the Great Moravian realm in the 9th century witnessed a considerable progress.
Its rulers accepted Christianity from Byzantium. The bloomery smelters worked on quite
large-scale using specific furnaces suitable for steelmaking and the blacksmiths learned
to apply advanced techniques involving iron-and-steel welding in various construction
schemes and heat treatment; a special attention deserve the welded-on steel edges on
the lyra-shaped strike-a-fires. More than one hundred metallographic analyses have been
carried out on iron tools. Nonetheless, the Slavs were more than willing to obtain weapons
and armour from the Carolingian West so that Charlemagne imposed an embargo and
issued five edicts blocking the export of these objects (..et ut arma et brunias non ducent
ad venundandum, for example).
The Great Moravian blacksmiths took a fundamental part in economic activity; they
produced stylized axe-shaped bars of four dimensions used as a premonetary currency. The
same task was undertaken by their counterparts in Scandinavia who forged not only axe-
shaped bars but also large amounts of spade- and scythe-shaped blanks. This economic
stage always preceded systems based on struck coins, and Scnadinavia and the western
Slavic world went through this stage during the 6th/7th to 9th centuries AD.
It seems that during the 10th - 12th centuries the work of blacksmiths became even and
standardized over the whole of Europe although analytical evidence is mainly available
from the eastern part of the Continent. The testing field might be seen in the cutlery.
Knives were constructed in several schemes of welding-on or welding-in of steel cutting-
edges (heat treatment like quenching or edge quenching and tempering became common).
The amount of all-iron blades has sunken to some 10%.
A specific chapter in the history of cutlery represents the production of early medieval
pattern-welded knives. Their cutting-edges were of steel, the central part was a pattern-
welded band and the blade back was of iron or combined with thin rods of phosphoric
metal. Natural etching created a permanent effect of light and dark zones. The manufac-
ture of these knives has been attested in England, Bohemia, Poland, Belarus and Russia
in towns, villages and, for example, in monastic settlements. The pattern rarely appears
as a copper-based inlay. The technique survived, in some sites, up to the 14th century.
The situation in Germany remains murky and southern Europe did not produced these
apparently expensive display and utility tools.
The smith’s tools of the early medieval period belong to the developed category. The
smithies are mostly known from rural sites and include both sunken-floored or free-
standing features and shelters. The iron objects appeared in settlement layers as lost
and abandoned things, in hoards, and in graves. Up to the 10th century warrior graves
were excavated with swords and sometimes with precious objects as well.

Blacksmith’s work in the High Middle Ages


The development of the hand-working blacksmith’s craft culminated in the 13th to 15th
centuries but simultaneously water-driven machine forges began to infiltrate.
Blacksmithing covered the most important technical material needs of everyday life to
an unprecendented degree. The number of kinds of iron ware including tools, weapons,
armour, utensils, devices and structural elements well exceeded a hundred. The smiths
living and working in towns enjoyed more freedom although all people were subjected
to feudal authorities (the king and the nobles). They were highly specialized: cutlers,
238 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
swordsmiths, armourers (some European cities like Nuremberg, Milan, Toledo and others
became celebrated by their armour production), farriers, locksmiths (who constructed
different locking mechanisms). Designers of tower horloges began to deliver their mas-
terpieces of clockwork. Many smiths stamped symbolic punchmarks (triangles, circles,
leaves etc.) onto their products; the right to use punchmarks could be inherited. Smiths
inhabited certain streets or quarters, often at the city gates. The social position of rural
smiths was incomparable with those in towns. They were fully subjected to feudal lords
and certain rules governed their work: apart of the production of some basic kinds of tools
and equipment they spent a lot of time repairing iron things, imported from the cities as
well (locks, special cutlery). A specific branch of highly specialized blacksmiths appeared:
they worked at silver and gold mines and daily maintained large numbers of mining picks
- tools equipped with steel points. Urban, suburban, rural and mine smithies have been
uncovered during archaeological excavations.
The technology of making tools lay in the application of standard techniques of welding-
on and welding-in of steel working parts, the time-consuming secondary carburization
essentially lost the importance for economical reasons. The locksmith applied the abrasive
precision cold work and soldering of the springing parts of lock systems. The hardness
of the steel elements was multiplied by quenching, or the partial quenching of cutting-
lines and tempering. The precise coldwork must have been applied in a newly introduced
branch of clock-makers; their horologes survived in many cases until present days.
The farriers also availed themselves with steeled iron better resisting the abrasion of
critical points. The armour manufacture still involved the making out of mail parts,
collars, hoods, shoulders but since the 14th century, plate armour used to be, step by
step, introduced. Steel breastplates and helmets rivetted together of several sheets of
mild steel might have been equipped with visors. Gradually, plate armour protected the
whole body. Large-scale production concerned the making of iron or steely crossbow-bolts.
There is lack of metallographic informantion on the manufacture of swords; apparently
they had steel blades and some examples show that their pommels could be made of cast
iron as well as some steel sheet for armourers.
Intentionally smelted cast iron began to be delivered from first charcoal blast furnaces
which were in operation above all in the Namur region during the 15th century (e.g. cast
iron shots are mentioned in the sources). Another branch of ironworking appeared on
the scene: the foundry. Indirect ironmaking took root in other regions in Europe, in the
Weald of Sussex and Forest of Dean in England and in Scandinavia as well (Lapphytan).
Far reaching processes can be shown on a specific category of worked iron: structural
clamps and rods used for reinforcing the ashlar mansonry of monumental architecture,
cathedrals and palaces. In the Palais du Papes at Avignon, the Grand Promenade con-
tained iron reinforcements produced by indirect process meanwhile the tested rods from
the Gallerie de Conclave and the Trouilles Tower were made of heterogeneous bloomery
material presumably worked in water-driven hammer-mills.
Hammer-mills operating with hydraulically moved hammerheads (themselves admirable
pieces of ironwork), together with the essential necessary anvils were able to process huge
blocks of iron, e.g. blooms from the so-called high bloomeries. The first hammer-mills
were already recorded in written sources in the 13th century (mollendinae ferri) but more
frequently in the 14th century (martinetus in France). An archaeologically published
hammer-mill combined with a bloomery ironworks was uncovered at Bargen (Switzer-
land, 15th century), yielding, for example, an oak beam from a hammer. There is no
CONCISE HISTORY OF EARLY EUROPEAN BLACKSMITHING 239
doubt that wooden machinery had started to dominate large-scale ironworking.

In postmedieval times the tasks of hand-working blacksmiths became more and more
restricted to meet local demands in villages. They forged some basic tools and repaired
damaged utensils. From the early 20th century they began to work with mineral coal,
used borax as flux to remove hammer scale and produced, as waste, amorphous dross
instead of PCB calottes. They were, however, skillful and often ranked amongst the local
philosophers.
Recently, the hand-working smith’s craft has turned to the production of artistic forg-
ings, specific domestic fireplace equipment and the like. Nevertheless, their work con-
stantly excites the interest of the public. Shows and performances demonstrating their
work are organized in many places and enjoy popularity. Hand-smithing schools still exist.
The commercional production of luxurious forgings applying, for example, the principles
of pattern-welding flourish in several European cities. However, the role of early European
master smiths was nowadays substituted by the industrial working of steel and the con-
sumption of the metal has reached astronomical amounts in machinery and construction.
By the way, everyone who walks through city streets is surrounded by iron sheet - the
automobiles.
GLOSSARY OF TECHNICAL TERMS

annealing, heating of steel under mild metal alloy in which two or more met-
temperatures of ca 600 - 650 ◦ C which als coexist in solid solution and pass
causes the recrystallization of the struc- from one crystal lattice to another (e.g.
tures. See Chapter V. face-centred, body-centred, cubic) and
austenite, a solid solution of carbon (and change in general properties.
other elements) in face-centred γ-iron, eutectoid steel, hard pearlitic steel of
see Chapters III, V, XI. about 0.86%C Steel with lower car-
bainite, one of the metastable structures bon content and a proportion of ferrite
of tempered steel, see Chapters V and is hypoeutectoid and that with higher
XI. carbon content and excess cementite is
billet, a well forged bloom suitable for hypereutectoid.
making a bar. fayalite, iron orthosilicate Fe2 SiO4 from
blank, here a kind of an unfinished (iron) the family of olivines, melting at about
artefact to be worked in further pro- 1180 ◦ C, is a constituent of iron and
cessing. many copper slags or occuring in vol-
bloom, a product of the direct smelting canic rocks (Fayal island in Azores). It
(bloomery) process. appears as dark grey laths on polished
bloomery process, the reduction of iron samples. In combination with other el-
ore under temperatures below 1500 ements (Ca, Mg, Mn) it forms different

C the result of which is molten iron mineralogical phases.
rich slag and a pasty iron sponge or ferrite, a ductile structure of iron con-
bloom. Occasionally small amounts of taining no or minimal content of iron
liquid cast iron were produced during a (0.02 - 0.03%C). See Chapters III and
bloomery smelt. XI.
carburization, introduction of carbon flux, a substance lowering the melting
into iron in the austenitic state. See point of metal and minerals. Sand flux
Chapters V and XI. (fine quartz) was used to remove ham-
cast iron, the product of blast furnaces, mer scale during the heating of iron in
smelted as liquid mass under conditions the smith’s hearth by melting to fay-
of the indirect process, containing more alitic slag.
than 2% of carbon. Grey cast iron con- glass, solidified amorphous substance
sists of pearlite and segregated flakes of also occuring (together with crystalline
graphite. Granulated graphite makes phases like wüstite and fayalite) in iron
the cast iron malleable. White cast iron slags. It appears in artefact inclusions
is characterized by cementite and lede- as well, containing e.g. wüstite some-
burite. To make steels, different cast times in form of minute dendrites.
iron sorts use to be heated in finery pro- graphite, in metallurgy carbon segregat-
cesses. ing during slow cooling of cast iron as
cementite, Fe3 C, hard and brittle iron microscopic flakes or granulated parti-
carbide with 6.67%C. See Chapters III cles. See Chapter XI.
and XI. gromps, from Polish gra̧pie, groats,
equilibrium phase diagram shows ef- minute particles of iron, extremely
fect of temperature and composition of heterogeneous in their carbon content
GLOSSARY OF TECHNICAL TERMS 241
which have been flaked from the re- ruling in the relevant region by the
heated iron sponge or bloom. weathering of volcanic rocks. It often
hammer scale, magnetic bluish-grey contains iron oxides. Greek nickelifer-
flakes of iron oxides (wüstite, mag- ous laterites are supposed to be used in
netite) formed on the surface of heated ancient iron smelting and their nickel
glowing iron. They could be knocked content lead to misunderstandings in
off or melted by sand flux (see Chapter terms of the use of meteoritic iron.
VII). Occasionally they occur as inclu- ledeburite, eutectic alloy of iron and car-
sions in the artefact bodies. bon (4.7%C) appearing in steels with
hearth, the lower part of furnaces or more than 1.7%C and in cast iron. See
smithing installation for heating iron. Chapters III and XI.
See Chapter VIII. magnetite, magnetic iron oxide Fe3 O4 in
hypereutectoid steel see eutectoid the spinel family of minerals. 1. A high
steel. grade iron ore. 2. The product of heat-
hypoeutectoid steel see eutectoid steel. ing of any ore in the process of roasting
inclusions, in metallography non- and the upper parts of the smelting fur-
metallic impurities in the crystalline naces under the effect of carbon diox-
metal matrix. In wrought iron and ide. 3. Constituent of hammer scale
steel they appear as deformed cavities and post-reduction slags.
filled with slag, often following the line martnesite, a metastable solid solu-
of welds or folds. Apart of silicates tion of carbon cooled rapidly from
and iron oxides, globular oxides and austenitic stage. See Chapters V and
sulphides may occur as well. XI.
ingot, in modern metallurgy a block of martquenching see quench-hardening.
cast iron for storage and subsequent metallography see Chapter XI.
processing. In archaeometallurgy mal- meteorites, cosmic mineral bodies im-
leable semi-products are denoted as in- pacting the Earth from interplane-
gots as well (bi-pyramidal ingots etc.). tary space. There are stony me-
iron, ferrum, Fe, a metallic element, teorites (aerolites, chondrites, achon-
group VIII of periodical system. See drites) and iron meteorites or siderites.
Chapter III. The latter are iron-nickel alloys (with
iron sponge, unworked spongy iron hav- some cobalt): kamacite contains 5 -
ing been sintered together in the slag 18%Ni, taenite 26 - 65%Ni. They
mass and around charcoal pieces in often reveal Widmannstätten texture.
front of the air-inlet in smaller and sim- Mesosiderites (pallasites) consist of
pler bloomery furnaces. During reheat- olivines embedded in iron-nickel cel-
ing and forging it was transformed to lules. Meteors which exploded and
more consolidated bloom or to black- did not penetrate to the Earth surface
smith’s starting stock at the cost of are called bolides. On potentional use
about 50% losses of the metal. of meteoritic iron by blacksmiths see
kamacite, nickeliferous α-iron in mete- Chapter I.
orites (5% - 18%Ni). The kamacite of Neumann bands, straight narrow lamel-
the 18th century BC Boldyrevo blade lae parallel to crystallographic planes
contained more than 9%Ni and more in the crystals of metal, mostly of iron,
than 0.6% Co. evoked by strokes in cold state or at
laterite, a residual soil formed under very low temperatures. They occur as
tropical climatical conditions, formerly twins.
242 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
nitridation, intentional saturation of sintering, in the bloomery smelting the
iron with nitrogenic substances. See coagulation of austenitic metal parti-
Chapter V. cles in the viscous slag to larger com-
PCB, plano-convex hearth bottom, a plexes (spongy iron), without being ar-
cake of smithing slag formed in tificially welded.
the charcoal-fed hearth, also SHB - sorbite, a formerly used term for a kind
smithing hearth bottom. See Chapter of fine pearlite in tempered steels. See
VII. Chapters V and XI.
phosphoric iron see phosphorus in iron. steel, a malleable alloy of iron, car-
phosphorus in iron. In the process bon and sometimes other constituent.
of bloomery smelting a part of phos- Nowaday all technical iron is denoted
phorus from ores (certaine haematites, as steel. In the past a hardenable
bog limonites) remains as P-pentoxide metal with more than 0.3 - 0.35%C. See
in the slag, another part transits into Chapters III, V and XI.
the metal, causing its brittleness (es- taenite, a solid solution of iron and nickel
pecially under cold conditions), con- (27 - 65% Ni occuring in sideritic me-
siderably increasing its hardness, not teorites).
worsening its weldability but hinter- tempering of steel, heat treating of
ing, to certain extent, its carburization. quench-hardened material below the
Phosphorus content above 0.1% is un- critical range (usually up to 500 - 600

wanted in modern steels. In prehis- C). Martensite transits to less brittle
toric and medieval times certain regions textures of fine pearlite. See Chapters
were directly dependent on phospho- V and XI.
ric iron supplied by ironworks smelt- tuyere, a nozzle protecting the mouth
ing phosphorus-rich ores and yielded of a blowing device. Blacksmiths used
iron with more than than 0.1% up to tubular or brick-shaped block tuyeres
1.0% P. Phosphoric iron was welcomed clay installed in the hearth. See Chap-
and sometimes intentionally used for its ters VIII and IX.
bright lustre. The Oberhoffer etching welding, the joining together pieces of
reveals phosphorus-rich zones on me- weldable metals at temperatures of 720
tallographic samples. Phosphorus also - 1200 ◦ C, here heated in a blacksmith’s
segregates in the welding seams (‘white hearth (forge- or fire-welding) and ham-
lines’, see welding seam). Phosphorus- mered. See Chapters V and XI.
rich ferrite grains reveal a grey moirée. welding seam, trace of joining of metal
See Chapters III and XI. pieces, here after forge-welding (contin-
quench-hardening, rapid cooling of uous chains of slag inclusions, segre-
steel from temperatures above 900 ◦ C gations of phosphorus, arsenic, nickel,
in water or more acid liquids in order to copper known as ‘white lines’, sharp
increase the hardness. It reveals brit- difference between crystalline struc-
tle acicular structure of martensite. See tures). The visibility depends on in-
Chapters V and XI. tensity and time of forging.
reheating of blooms, heating of iron Widmannstätten structure, a
sponges or blooms regularly in special triangle-shaped texture of carbon
hearths in order to remove as much ad- steels, developed at high temperatures
hering slag as possible. About a half of ca 1000 ◦ C from austenite and ferrite
of metal oxidizes and considerable loss along the split-planes of γ-iron and rel-
involves the process. atively rapidly cooled, e.g. in air (not
GLOSSARY OF TECHNICAL TERMS 243
quenched). It appears in nickel-iron al- phase in the reduction of iron ore; in
loys of meteorites as well. See Chapters can contain some MnO or NiO2 . Com-
III and XI. bined with SiO2 it becomes a con-
wrought iron, a malleable low-carbon stituent of fayalite. Excess wüstite ap-
iron with much entrapped slag pears in bloomery slags and slag inclu-
(bloomery iron, in post-medieval met- sions in iron artefacts. Together with
allurgy fined cast metal as a material magnetite it occurs in hammer scale.
for subsequent smithing and rolling). The name is in honorem Fr. Wüst, a
wüstite, iron monoxide FeO (Fen O), a German metallurgist.
GLOSSARY OF HISTORICAL AND ARCHAEOLOGICAL
TERMS

Achaemenids, Persian dynasty which Charlemagne at the end of the 8th cen-
derived itself from Achaemenus, the tury AD. Surviving communities were
great-grandfather of Cyrus the Great rapidly assimilated by Slavs. Among
(559 - 530 BC), the founder of the Per- Avars were skillful arrow- and sabre-
sian relam. He was buried in Pasar- smiths.
gadae. The dynasty died out by Xerxes Balts, Indo-Europeans inhabiting the
II 424 BC. lands of the eastern Baltic sea area.
Aeschylus (ca 525 - 456 BC), a tragic Their language (Lithuanian, Lettish,
poet who in his Prometheus desmotes Prussian) shows considerable affinities
(Prometheus Bound) mentions the to the early Slav language; Slavs were
quenching of iron/steel and the iron- neighbours of the Balts in the south.
making Chalybes. Some linguists presuppose a prehistoric
Alesia, Mt. Auxois, now Alise-Ste-Reine Balto-Slavic unity.
in Côte d’Or, Burgundy, a Celtic op- BG, Commentarii de bello Gallico, Cae-
pidum and the place of the last resis- sar’s report on the conquest of Gaul 58
tance of Gauls against Rome. Excava- - 51 BC. C. I. Caesar (100 - 44 BC),
tions started under Napoleon III in the one of the triumvirs, was the procon-
19th century. sul in Cisalpine and Narbonensis Gaul
Amelungenlied, a poem celebrating the and commander of troops garrisoned
deeds of the Gothic house of Amali or there so that he was able to start the
Amelung of the 6th century AD. The campaigne. The Commentarii include
preserved text comes from later cen- remarks on life, culture and religion
turies and involves the legend of Wey- of Celts and some Germanic tribes.
land the Smith and his art of sword- He touched iron several times: Celtic
smithing. iron mines, iron production in southern
Anglo-Saxon period in Britain covers Britain, taleae ferreae.
the time 5th to 11th centuries AD. Im- Bibracte, the hill of Mont Beuvray,
migrated Angles, Saxons and Jutes set- a Celtic oppidum in the French de-
tled in Britain during the 5th century partement Saône-et-Loire, the centre
AD (Pagan period, 5th - 6th centuries of the Haedui tribe, denoted by Cae-
AD). After the conversion to Christian- sar as the most flourishing an impor-
ity (St. Augustin, AD 506) England tant in the entire Gaul. Large-scale
was established. Scandinavian raids excavations dated from 1865 to 1895.
took place up to the Norman conquest The Come-Chaudron valley housed ar-
in AD 1066. tisanal workshops, smithies as well. Af-
Avars, Asiatic nomadic tribes of Turkish- ter 50 BC, the successor of Bibracte be-
Tatar origin. In central Europe they came the town of Augustodunum (Au-
appeared in the mid-6th century AD tun).
and came into contacts with early Biringuccio, Vannocio (1480 - 1539),
Slavs. In Pannonia they ruled from AD the author of the treatise ‘De la
568. The Avar realm was destroyed by pirotechnia libri X’, Venice 1540, edited
GLOSSARY OF HISTORICAL AND ARCHAEOLOGICAL TERMS 245
one year after his death. Iron is treated northern part of the country. Crema-
in book I 6 and IV 3. He mentioned tions as well as inhumations belonged
specialized smithies at mines where to the funeral rite. A great deal of pot-
miner’s tools were repaired. During his tery is very fine, fired-red with black
dramatic life, Biringuccio was the di- painted geometric motifs (certain kinds
rector of the iron mines and, in his last of pottery was made exclusively for
years, the head of the papal foundry burial). The chieftans of that folk (pre-
and armoury. supposed to have been proto-Celtic)
Bronze Age, the archaeological term for were buried in large chamber graves (ca
period of human culture characterized 5m x 4m), richly equipped not only
by the use of weapons and tools made with pottery (up to 60 vessels in one
of bronze - a copper-based alloy with princely grave) but also with bronze or-
tin or arsenic. Considered in time and naments and iron sword and knives of
space, the term is relative. In the Near the chief. Four-wheeled wagons or their
East the Bronze Age (involving his- parts and splendid bronze-fitted horse
torical time and advanced civilization) yokes belonged to the grave gifts. Until
ends about 1200 BC, whilst in central now, settlement traces are rare.
Europe it was prehistoric and contin- Carthusians, a monastic order founded
ued up to ca 800 BC (Reinecke H B3 ). by St. Bruno at Chartreuse near
In Scandinavia, the latest stages are Grenoble, eastern France, in AD 1084.
considered to last until 500 BC (Mon- The Carthusians were engaged in hand-
telius VI). During the Bronze Age, the icrafts and also in iron mining and
first isolated iron objects appeared in working.
wide areas. Celts, the western-most Indo-European
Býčı́ skála (Bull’s cave), Moravian population in Europe, the ethnogen-
Karst, a sacrifice place or princely esis of which has had to be com-
burial of the Hallstatt D period (5th pleted at the decline of the Bronze Age
century BC) comprising two large fired (Upper Danube tumuli and urnfields).
areas with remains of wagons with iron- (Keltiké is mentioned by Hecataeus
tyred wheels, bronze vessels and bronze and Herodotus mentioned the Celts
and gold jewelry. Ritually sacrified hu- as well). From central Europe they
man beings (more than 40 individu- expanded westwards, colonized Gaul
als, mostly women) and animals were (Galli), parts of British Isles, the north
found. Burnt grain appeared in layers. of Spain (Celtiberrians). Later, they
In the rear part of the cave a work- raided and colonized northern Italy
place called ‘smithy’ has been uncov- Gallia Cisalpina, Balkans, Carpathians
ered (heavy iron sledge hammers, a half and Asia Minor (Galati). Most of these
of a bipyramidal ingot, smithing slag). regions they left during the 1st century
After a find of a bronze bull statuette BC or were assimilated.
with iron inlays near the entrance of the Chalybes, an ancient tribe in the western
cave, J. Wankel decided in 1870’s to ex- Pontic area of Asia Minor, legendary
cavate the layers of the cave complex ironworkers (sidérotektones by Aeschy-
and published his discoveries in 1882. lus, 525 - 485 BC; Xenofon mentioned
Bylany culture, after Bylany in east- iron ore mining Chalybes when lead-
ern central Bohemia, a late Hallstatt ing the Greek mercenaries during the
period culture of the 7th to the 5th Anabasis, 401 BC). The tradition must
century BC, spread in the central and be old since at that time chalyps be-
246 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
came a notion for steel, also in later according to the Nordic legend, Colchis
periods. Chalybes possibly washed rich was inhabited by dwarfs who learned
magnetite sands of their area which was Weyland the Smith to his art.
inhabited, according to Xenofon, by the Cypriot and Cyprogeometric peri-
Colchoi which alludes to the near-by ods. The island of Cyprus in the
Colchis with its iron production around eastern Mediterranean is known, above
1000 BC. all, by its ancient copper production,
Cistercians, members of a monastic or- already during the local Bronze Age
der of Benedictine rule, founded by called Cypriot period. In the late phase
St. Robert at Cı̂teaux (Cistercium) Cypriot III (12th century BC) first rare
in France AD 1098). The Cistercian iron knives appeared. Especially iron
monastery at Clairvaux (Clara Val- knives came to light in the subsequent
lis) became the progenitor of many Cyprogeometric period in which the is-
daughter monasteries in central Euro- land saw more intensive contacts with
pean countries (12th - 14 th centuries ancient Greece. At that time, Cyprus
AD). The monks were much engaged in entered the Early Iron Age: CG I (late
handicrafts, especially in making and 11th century), CG II A, B (10th cen-
working of iron. tury BC).
Cimmerians, tribes of semi-nomadic Dacians, Daci, the Roman name for
Indo-European branch of Iranian ori- the southern Thracians (Greek Getes)
gin, possibly related to Thracians, men- in Romania, part of Hungary and
tioned by Homer, Neo-Assyrian epigra- eastern Slovakia. It was an Indo-
phy and Herodotus. From north-Pontic European population already men-
steppes they advanced to Carpathi- tioned by Herodotus. Its material cul-
ans (Thraco-Cimmerian style) and, in ture was similar to that of the Celts.
the 8th century BC, along the eastern The capital was Sarmizegetusa. Dacia
Black Sea coast to Asia Minor where was a Roman province between AD 106
they founded a kingdom. This was de- to 275.
stroyed by the Lydians in the early 6th Damascus steel, damascene steel (fulad
century BC. In the Ukrainian steppes in Persian, bulat in Russian), a hyper-
they were subordinates to Scythians eutectoid hard carbon steel which was
(7th - 6th centuries BC). According traded through Damascus in Syria from
to some viewas they should have con- at least 6th century AD. Damascene
tributed to the spread of certain iron swords and sabres were produced up to
artefacts to central Europe. the 18th century, especially in Persia.
Classical period in ancient Greece, the They are characterized by their wave-
5th to the 4th century BC. like pattern moiré revealed, after etch-
Colchis, a historical territory of the ing, by the dispersion of globular ce-
south-western promontories of Cauca- mentite in the pearlitic matrix. Ex-
sus (part of modern Georgia) and ad- perimental work by Anosov, Belayev,
jacent to the Black Sea coast. The Johansen and others was successful in
gold-bearing rivers gave rise to the an- making some replicas but the original
cient legend of the Golden Fleece of the technique remais veiled.
Argonauts and forests supplied wood Demosthenes (384 - 322 BC), the great-
for ship building. In addition, impor- est orator of antiquity. In his speach
tant iron production has been discov- against Aphobus he mentions a weapon
ered (around 1000 BC). Interestingly, manufactory owned by his father bor-
GLOSSARY OF HISTORICAL AND ARCHAEOLOGICAL TERMS 247
ing the same name. Specialized slave to Germanic folks in the English us-
blacksmiths produced weapon blades in age) invaded Gaul in the late 2nd cen-
that workshop. tury BC. After the breakdown of the
Etruscans, Lat. Tusci, Greek Tyrrhenoi, Celtic power one century later, Ger-
Rasena in their own language, a non- manic groups conquered territories up
Indo-European population in west- to Rhine and Danube and became
ern Italy (Tuscany) and subsequently neighbours of Rome. Subsequently,
northwards. Their origins are dis- they widened their expansion to the
cussed. According to Herodotus and east (see Goths).
the Lemnos Inscription they sailed from Gortyna laws. In the ruins of the town
Lybia and settled in coastal western of Gortys in southern Crete there were
Italy in the early 1st millennium BC found, in 1884, inscriptions concerning
(despite the view of Dionysius of Ha- the laws from the period around 450
likacarnassus). Their language, as BC. They are the earliest documents of
a whole, is not deciphered although the Greek civil law. Iron tripods in the
their script derived from that of the role of currency bars are mentioned.
Greeks. Etruscans have played a con- Goths, a Germanic tribal union the
siderable role in smelting a working ethnogenesis of which developed in
iron. Their religion-based confedera- southern Sweden (Vestergötland, Got-
tion of 12 towns saw its decline in the land). Goths expanded to the Vistula
5th century BC. river estuary and in the 2nd century
fanum, in the Roman culture a sacred AD they moved south- and eastwards
place. to modern Ukraine. Possibly they par-
Geometric period and style in ancient ticipated in the Tchernyakhov culture
Greece (9th - 8th centuries BC). The (2nd - 4th centuries AD) which also in-
geometric ornaments on pottery were volved Dacians, Sarmatians and earli-
enriched by meanders. At that time, est Slavs. Subsequently, they were di-
Greece entered the fully-fledged Iron vided as Ostrogoths and Visigoths.
Age. Ostrogoths (eastern Goths) ruled the
Germania, the Latin name for territo- territories from the Black Sea up to
ries inhabited by Germanic (Teutonic) the Baltic under their king Ermanarich
tribes. Parts of Germania became Ro- (4th century AD). Huns destroyed their
man provinces (Germania Superior on realm. Ostrogoths settled in Pannonia
the left bank of middle Rhine and Ger- and under Theoderich they overruled
mania Inferior between Ardennes and Italy. Visigoths (Western Goths) in-
the lower Rhine). Except of the Agri habited, during the reign of Constan-
Decumates (now Württemberg), the tine the Great (4th century AD) the
rest of central Europe was Germania northern lands of the Danube estuary.
Libera (as to Tacitus) or Megalé Ger- Later, under Alarich, they attacked
mania as to Ptolemy of Alexandria. Italy and conquered Rome. However,
Germanic (Teutonic) tribes, an Indo- as allies of Romans, they fought against
European population which settled, Huns.
presumably from Neolithic times, Scan- Great Moravia, in Byzantine sources
dinavia and northern territories of cen- Megalé Moravia. The first attempt of
tral Europe. Gradually they ex- the establishment of a realm among
panded west- and southwards. Cim- western Slavs in the 9th century AD.
bri and Teutoni (the latter gave name Having a good economic background,
248 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
but in the conflicts with The East tite capital.
Franconian empire, it spread in mod- Hecataeus of Miletus, a Greek historian
ern Moravia, southern Slovakia and fi- (logographos), 550 - 476 BC, predeces-
nally Bohemia and south Poland (Cra- sor of Herodotus. He produced a de-
cow). The rulers (duces, reges) of the scription of the Earth and mentions
Mojmı́r dynasty are known by name. that northwards of the Liguri a land
In 863/64 Christianity of the Byzantine begins which is called Keltiké.
type was introduced by Constantine Helgö, a small island in the Mälaren lake
and Methodius from Thessaloniki. The system east of Stockholm, with a com-
invasion of Magyars (early Hungarians) plex of sites having been flourishing in
caused the fall of Great Moravia in the AD 400 - 800, a centre of the beares
early 10th century AD. of the Vendel culture, predecessor of
Haithabu, Hedeby, a Viking period for- that of Vikings. Large-scale excava-
tified centre and trading post with har- tions took place since 1950’s (house
bour at the Schlei estuary in Schleswig, plans, metallurgic workshops, smithies,
southern Jutland (9th to 11th centuries rich find inventory).
AD). Connections with Scandinavia. Helladic culture and period, Bronze Age
Handicraft activity including the black- in the continental Greece. Three main
smith’s work attested by archaeologi- periods: Early Helladic 2055 - 1900 BC,
cal sources (amounts of smithing slag). Middle Helladic 1900 - 1600/1500 BC,
A long wall across Jutland, the Dane- Late Helladic (1500 - 1100 BC). The
werk, joins the western fortification of latter phase saw the beginning of the
Haithabu. active ironworking. The successor was
Hallstatt, an archaeological site in a val- the Mycaenean culture.
ley above the town and lake of Hall- Hellenistic period, historical phase from
statt in Upper Austria. A large and 325 BC up to BC/AD. Greek, Asia
richly equipped cremation and inhu- Minor and Egyptian cultures yielded a
mation cemetery (some 1300 graves, syncretism. One of the leading centres
26% of which had iron weapons). The was Alexandria.
cemetery dates to the period between Hephaestus, Héfaistos, originally a de-
800 and 500 BC, in several phases. mon of fire (Asia Minor), in Home-
The prosperous community controlled ric epics the god of fire and skilled
the nearby underground salt mines and smith having produced the armour of
traded with salt over long distances. Achilleus etc. The centre of his cult was
The site is eponymous for the Hallstatt the Lemnos island, in Athens was the
period, culture, style. temple Héfaisteion. In Rome the role
Hallstatt period comprised, after Rei- of the god of smiths played Vulcanus.
necke, early stages HA - HB (Late Herodotus of Halicarnassus, a Greek his-
Bronze Age) followed by Early Iron Age torian of the 5th century BC (‘Father of
(HC - HD, 7th to 5th century BC). The historiography’) who compiled the His-
periods are now subdivided (e.g. HB 3 torié apodexis giving account not only
means 9th/8th century BC, with early on Persian wars but also of non-Greek
occurence of iron artefacts). (barbarian) folks in east and west. In
Hattians, non-Indoeuropean predeces- IV.49 he names the Celts at the springs
sors of Hittites in Asia Minor. Their of Istros-Danube.
capital might have been already Hatti Hesiod, a Greek rhapsode and poet of
- Hattusass - Boghaz-köy, the late Hit- the 8th/7th century BC having lived
GLOSSARY OF HISTORICAL AND ARCHAEOLOGICAL TERMS 249
and worked in Askra, a Boeotian vil- hoards; it is unevitable to consider the
lage. Author of Works and Days (Erga find circumstances in individual cases.
kai hémérai, blacksmithing mentioned) Holy Cross Mountains (Góry
and Theogony-Birth of Gods (Iron Age Świȩtokrzyskie in Poland, an isolated
mentioned). mountain ridge near Kielce. A no-
Hippocrates of Cos (460 - 377 BC), the tion for its La Tène and Romano-
most renowned physician of antiquity. Barbarian iron production based on
His 58 treatises are collected in the good haematite from an underground
so-called Corpus Hippocraticum. He mine. More than 100 excavated sites
mentions the tempering of steel in oil concentrated on ca 100km2 .
(sidéros baféntos eis elaion). Homeric epics deal with the siege and
Hittites, Indo-European population oc- conquest of Troy or Ilion in western
cupying Asia Minor after about 1900th Asia Minor (Iliad and with the wan-
century BC, successor of non-Indo- dering and home-coming of one of the
European Hattians. Hittites left their war participants - Odysseus Odyssey).
homeland somewhere north of Cauca- The events were probably touched by
sus and Caspian Sea. About 1500 aoids and rhapsodes, ancient minstrels,
BC their rulers advanced in creation of during the 10th - 9th centuries BC but
a realm which controlled adjacent re- were epically transferred to two leg-
gions. The deeds and culture are illus- ends which are attributed to Homer.
trated in cuneiform texts from archives. The personality of this poet remains
They are written in an Indo-European veiled but despite of some 19th century
language and contain many references doubts the prevailing view respects him
concerning iron and ironworking. The and admit that he had lived somewhere
capital of the Hittite confederation in the western Asia Minor. Pisistrates,
was Hattussa, now Boghaz-köy, on the the tyrant of Athens in the 6th century
Halys river. The fall of the Hittite BC, is held for having been responsible
realm dates from about 1200 BC. for the first complete redaction of both
hoard, a complex of collected goods poems. There are 24 references to iron
buried on limited spot in a pit, in the Ilias and 25 in Odyssey: they
bag, chest, covered by stones etc. show a ‘stratigraphy’ - from places re-
The reason of hiding was different sembling the rarity of iron in the late
sothat in European languages a lot of Bronze Age up to the famous descrip-
terms appeared: German Sammelfund, tion of steel quenching (Od. IX 391).
Hortfund, Massenfund, Depotfund, in Iron contributes to the statement that
French dépôt, in Russian klad, in Pol- the topic of the legends originates in
ish zespól in Czech hromadný nález. different periods and the final version
Should it consist of precious things or of must have been completed in the full
coins the term treasury is equivalent Iron Age, somewhen in the 8th century
to German Schatzfund, Polish skarb or BC.
Czech poklad. Some hoards were hid- Illyrians, an Indo-European population
den just to be used again, other should which, since the 2nd millennium BC,
be buried forever: to be blocked and setled the north-western part of the
kept out of circulation or left as a votive Balkan penninsula. In the north-
gift (those spread over vide areas see ern Italy their neigbours were related
sacrified deposit). It is uneasy to inter- Veneti. The craftsmen of both groups
pret the factual reason of the burying of played a role in transmitting experi-
250 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
ence concerning the ironmaking and - La Tène period group named after an
working. incineration cemetery with fine pottery
imbrex, pl. imbrices, Latin term for fired ware, presumably burying strongly Cel-
ridge roof tiles. Tegulae are bricks. tized Germanic newcomers. Numerous
Indo-Europeans, the greatly spread iron objects (weapons, personal gear
population family the genesis and ex- etc.) as grave gifts.
istence of which were presented exclu- Kyjatice culture, named after a cre-
sively by linguistics. The primeval lan- mation cemetery in S Slovakia, a part
guage of Indo-Europeans was recon- of Hungarian-Slovakian Piliny culture
structed on the basis of terms being complex of the Late Bronze Age (HB).
common in various historical and mod- Rare iron objects of the 8th century
ern languages. These terms indicate BC.
that the original homeland of Indo- La Tène, an archaeological site on
Europeans was a great part of Asia and the eastern shore of Lake Neuchâtel,
eastern Europe with plains (suitable for Switzerland, the territory of Celtic Hel-
nomadic cattle-breeding) and moun- vetii. During land development in the
tains with snow etc. Subsequently, late 19th century, there was discovered,
presumably during the 3rd millennium in swamps, an enormous deposit of iron
BC) they split into many branches ex- objects (weapons, tools), wooden and
panding to all geographical directions leather artefacts concentrated around
and their langue differentiated. With two timber jetties protruding into the
exception of Finno-Ugrians, the Indo- lake. This indicates an important sac-
Europeans occuppied the entire Europe rifice place (about 100 BC). The site
and vast parts of southern Asia (Irani- of La Tène became eponymous for the
ans, Aryans in India). late pre-Christian millennium, ancient
Iron Age. As obsolete has to be classi- Celtic culture and style.
fied the notion when used for indication La Tène period, in continental Europe
that in the human culture first iron ob- time between about 500 BC - O, char-
jects appeared. Nowaday, a functional acterized by Celtic material culture as
approach has to be accepted for indi- revealed by archaeology. Tischler dis-
vidual geographical areas and chrono- tinguished three periods: I (400 - 300
logical periods. See Reinecke, Mon- BC), II (300 - 100 BC), III (1st cen-
telius, and here Chapter II. tury BC). This classification is still in
Khazars, an ancient Turkish population use in France. Reinecke elaborated an-
which spread south-westwards from other system for central Europe (L A to
Ural and played a dominant role in D). The first precedes Tischler I (i. e.
southern Russia (7th - 9th centuries 500 - 400 BC). Numerous scholars have
AD). During subsequent two centuries contributed to more detailed subdivi-
they were mostly assimilated by the sion of individual phases (e. g. LC1 ,
Slavs. The Saltovo-Mayatsk culture in LD2 etc.).
southern Russia and Ukraine is con- Limes Romanus, the fortification sys-
nected with Khazars (9th century AD). tem of the Roman Empire border (1st
Khorsabad, ancient Dur Sharukin in N to 4th centuries AD). The spin of the
Iraq with Palace of Sargon II (771 - 725 European limes was military roads con-
BC); the Victor Place’s discovery of a necting the bases like burgi, castra,
huge store of iron, see Chapter IV. castella along Rhine and Danube and
Kobyly culture, a north Bohemia Late across the land between Koblenz and
GLOSSARY OF HISTORICAL AND ARCHAEOLOGICAL TERMS 251
Kelheim in Germany (Agri decumates, of ironwork of every kind, unique epi-
abandoned AD 260). In Scotland there graphic documents concerning the ex-
was the Antonine wall and in north- port of Norician iron over the Roman
ern England the Hadrian’s wall (both world.
abandoned after AD 400). On the mid- Magyars, early Hungarians, a Finno-
dle Danube the limes was protected Ugrian branch of a nomadic population
by bumper zone of territories in mod- which moved from Russian territories
ern southern Moravia and SW Slo- (Volga, Oka, Kama rivers) down south
vakia. The Dacian-Moesian limes de- and separated during the 5th century
fended the lower Danube. During the AD. Later, in the 9th century, they in-
flourishing periods of the Empire the vaded the Carpathian basin and settled
limes divided thr advanced Roman civ- in Pannonia. Shortly after AD 900 they
ilization from the rural ‘barbaricum’. destroyed the Slavic Great Moravia and
During the 4th and 5the centuries AD established their own state, Hungary.
it was penetrated by Germanic and Manching, an important Celtic settle-
Sarmatian tribes and gradually ceased ment centre near Ingolstadt, Bavaria,
to be defended. Germany. During the 3rd - 2nd cen-
Llyn Cerrig Bach, a Celtic sacrifice turies BC intensive land covering, since
place on the island of Anglesey (Mona) the late 2nd century a fortified op-
in England. Iron weapons, wagon pidum (the eastern-most murus Galli-
tyres, fetters etc. having been thrown cus). Sanctuaries, estates, workshops,
into a swamp (Early Iron Age, last cen- rich traces of ironwork.
turies BC). Published by Sir C. Fox in Mästermyr, a moor in southern Gotland,
1946. Sweden. In 1936 a hoard of black-
Lusatian culture, named by the Ger- smith’s and carpenter’s tools and other
man physician R. Virchow who rec- irons (e. g. bars) has been discovered
ognized the finds in Lower Lusatia. in an iron-mounted wooden chest. The
A cultural complex with many local reason of burying is not clear, a votive
groups dating from the Late Bronze gift as well as a transport accident of a
Age up to the Early Iron Age and craftsman are taken into account.
covering the territories between the Migration period, in Europe the time
Elbe river and both river sides of Vis- of the late 4th up to 6th/7th centuries
tula and including Moravia and eastern AD). The events started with the severe
part of Bohemia as well. The funeral impact of Huns from the steppes in the
rite was cremation (Urnfields). Settle- East (Attila reached Gaul in AD 451).
ments with free-standing and sunken- Many Germanic tribes (Visigoths, Van-
floored houses, fortified centres, devel- dali, Alamani, Burgundi etc., and later
oped metallurgy of bronze, rare small Longobardi and Franks) left their home
iron artefacts (rings) already in the regions and moved having destroyed
Late Bronze Age. The bearers of the Roman limes. Angli, Saxons and
the Lusatian culture might have be- Juts crossed the sea for Britain and left
longed to the Veneti branch of Indo- their settlements in north-western Eu-
Europeans. rope. New ethnics (Slavs, Avars, Mag-
Magdalensberg, originally a Celtic yars) arrived as colonists or invaders to
(Norician) oppidum in Carinthia, Aus- central Europe.
tria, subsequently a Roman town of Minoan period after the mythical king
the 1st century BC. A rich collection Minos in Crete dates from the Early
252 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Bronze Age (3rd millenium BC), dur- upper and middle streams of Euphrat
ing 2100 to 1600 BC first palaces ap- and Tigris rivers inhabited by the
peared on the island. The Late Minoan Semitic Assyrians. After the early pe-
saw the rare application of iron (gold- riod in the late second millennium BC a
plated iron finger-rings). The term Mi- powerful realm was created during the
noan culture has been introduced by Sir 9th century BC by Assyrian rulers. It
Arthur Evans. was destroyed by Medes and Persians in
Mitanni, an ancient realm between the 618 BC. The Neo-Assyrian kings took
upper Euphrat and Tigris rivers (since enormous spoils, taxes and tributes in
the 16th century BC), a neighbour of iron from all of the Near-Eastern king-
the Hittite confederation. In the 14th doms and city-states.
century BC the king Tushratta main- Noricum, a Celto-Illyrian kingdom (Reg-
tained contacts with Egypt and tried num Noricum) on the territory of mod-
to marry out his daughter to pharao ern Austria, having been established in
Amenophis IV. Among the bridal gifts the 2nd century BC. During the reign
are named, according to tablets from of emperor Augustus it became a Ro-
the Tell el Amarna archive, amulets man province. In the history of iron the
and blades of habalkinnu, iron or steel. renowned Norician steel was exported
Montelius, Oscar (1843 - 1921), a lead- to the Roman world (see Magdalens-
ing Swedish archaeologist, the promo- berg and Chapters VIII and IX.
tor of Scandinavian prehistoric typol- Notitia dignitatum, a list of Roman
ogy and chronology based on the evo- officers and civil and military magis-
lution of bronze artefacts. Periods M trates (late 4th century AD). It con-
V and especially M VI (7th - 6th cen- tains, among other data, information of
tury BC) represent the coming of iron Roman large-scale fabricae having pro-
to Scandinavia. European Reinecke duced iron arms and armour in different
chronology: HB3 , HC, HD. parts of the Imperium.
murus Gallicus, Gallic rampart, de- oppidum, a Latin term for fortitified set-
scribed by Caesar in the mid-1st cen- tlement (munitium opis causa) having
tury BC (BG VII, 23). A structure of been used by Casar generally for Celtic
wooden beams crossed in right angle fortified centres. In archaeology, Celtic
which were put in distances of about oppida occur during the 2nd/1st cen-
60cm (2 Roman feet). Supported by tury BC and yielded magnificent mate-
stones, the next levels were gradually rials concerning the spiritual and mate-
superimposed until the desired height rial culture (France, Bavaria, Bohemia,
of the wall was reached. Timbers pro- Moravia, Central Germany). Later,
truded to the outer side which was apart of some buildings in front of Ro-
faced with worked ashlar stones. What man circuses, the term oppidum de-
was revealed by archaeology was that noted any medieval little town as well.
the timber crossings were joined by Ostrogoths see Goths.
large iron nails (15cm - 20cm). This Penkiv’ka culture, 5th -7th century
type of rampart was built in Gaul and AD, in the Ukraine, characterized
in Bavaria up to Manching. Kilometers by open agricultural settlements with
of such fortifications required a hardly sunken-floored houses and rare hill-
imaginable number of tonnes of struc- forts (Pastyrskoye gorodichtche, with
tural iron. numerous iron artefcts). Ironworks at
Neo-Assyrian realm, territory between Hayvoron. Incineration and inhuma-
GLOSSARY OF HISTORICAL AND ARCHAEOLOGICAL TERMS 253
tion burials. The bearers are supposed part of Central Germany, Moravia, Slo-
to be ancient Slavic Antes or Antai. vakia up to Ukraine (Pripet, Prut and
Phoenicians, a Semitic population hav- Dniester river valleys) incl. the vari-
ing inhabited the coast of the east- ant called Korchak type). Excava-
ern Mediterranean since the 2nd mil- tions have shown rural settlements with
lennium BC. Gradually Phoenicians sunken-floored houses, granary pits etc.
founded urban states and important The burial rite was cremation. In
ports as trading posts (Ugarit, Tyrus, the east, in tracts around the Dnieper
Byblos, Sidon etc.). About 1000 river the early Slavs are connected with
BC they began to colonize the West: the related Pen’kivka culture (6th -
Iberian coast and northern Africa 7th centuries AD). As to ironworking,
(Hadrumetum, Carthage). As skilled rather simple techniques in the blade
sea-farers they ruled the Mediter- construction prevailed in early phases
ranean. They are held for authors of of ancient Slavic cultures.
the script having been the model for Protogeometric pottery and period in
the Greek alphabet. ancient Greece of the late 11th to
Pliny the Elder. C. Plinius Secun- 10th centuries BC, characterized by ce-
dus Maior, (AD 23 - 79) a high posi- ramic ware with circular, semicircular
tioned Roman official, the author of the or rhombic ornaments and triangles.
encyclopaedia Historia naturalis in 37 The beginnings of the functional use of
books. As a naval commander he lost iron.
his life when observing the eruption of Przeworsk culture, after the Romano-
Vesuvius in AD 79. Barbarian cremation cemetery in
Pliska, near Shumen, Bulgaria, ruins southern Poland. The culture involves
of the first capital of the Bulgar- the lands of central, eastern and above
ian realm (in AD 681 - 893) include all southern Poland in the Late La
the citadel, fortifications, town with Tène and Romano-Barbarian periods.
churches, workshops incl. smithies The amber route passed through its
and a clay water-tube system. Czar territory and large-scale pottery work-
Symeon transferred the capital to shops, influenced by Roman technol-
Preslav (AD 893), about 50km south- ogy worked near the modern Cracow.
wards. The cultural province involved the Holy
Pompeii, an ancient town on the western Cross Mountains smelting region, the
coast of Italy. inhabited by Osci, Etr- leading one in the barbaricum, as well
uscans, Samnites, later a centre of lux- as the ironworks in Masovia west of
urious sites of the Roman upper class. modern Warsaw.
Destroyed by the eruption of Vesuvius Reinecke, Paul (1842 - 1958), an out-
in AD 79. Excavations since 1718. standing German archaeologist, the
Many traces of artisanal workshops and founder of Central European chrono-
shops, incl. smithies. logical system covering the Bronze Age,
Prague culture, originally denoted as Hallstatt and La Tène periods each
Prague type of early Slavic culture as comprising phases from A to D. These
described by I. Borkovský in 1940 on were later subdivided by many authors.
hand of cremation cemeteries on the The system is in use until the present
territory of Prague with typical hand- day. Less known is that Reinecke un-
made tall pots. Subsequently its spread dertook a large-scale mapping of iron
had been recognized over Bohemia, ore and bloomery smelting places in
254 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Bavaria (published 1936). Sarmizegetusa, now Gradiştea
Romano-Barbarian period, in non- Muncelului in Romania, the principal
Roman Europe the time-span of the Dacian fortified site, 1st century BC to
first four centuries AD, corresponding 1st century AD (conquered by Romans
with the Roman Empire (in German AD 105/106). Sanctuaries, workshops,
römische Kaiserzeit, RKZ). This was hoards of iron objects.
the time of the rural civilization of the sax, scramasax, a one-edged short sword
Germanic tribes in the Germania Lib- with flat handle used by Germanic war-
era. riors during the Migration period, fre-
sacrified deposit, a complex of goods quent burial gift of men. Examples
(precious objects, metal artefacts etc.) longer than 50cm are rare (in German
having been set out of circulation and Langsax). Numerous saxes and scra-
buried not to be used again. The mat- masaxes were equipped with pattern-
ter is not only of certain hoards but es- welded parts and serrated welds joining
pecially of large deposits in sanctuaries the steel cutting-edge.
(ancient Greece, Celtic Gournay-sur- Scythians, a complex of nomadic Indo-
Aronde, France), in moors and rivers European (Iranian) tribes who lived
(Celtic La Tène, Llyn Cerrig Bach in from the 8th century BC in steppes and
England, Romano-Barbarian Vimose, forest-steppes north of the Black Sea
Illerup, Thorsberg in Scandinavia). In and Caucasus up to Altai. They raided
all of them a quantity of iron artefacts neighbouring countries. The huge
was found, especially swords and other burial mounds (kurgans) of Scythian
weapons. chieftains were magnificently equipped
Saltovo-Mayatsk culture, named af- with luxurious goods. During the
ter fortified sites and cemeteries 4th century BC the so-called ‘Scythian
at Verkhne Saltovo and Mayatsk, kingdom’ was established on the lower
Ukraine, is ascribed to Khazars who, Dnieper river. Scythians were out-
having been driven by Huns, settled standing horsemen and archers, used
between the Upper Donec and Don two-wheeled as well as four-wheeled ve-
rivers and in adjacent regions. They hicles. Skilled metalworkers, black-
reached much success in developing the smiths and armourers worked among
ironworking techniques. Many shapes them.
of objects, e.g. axeheads, were adopted Slavs, a mighty ethnic branch of Indo-
by Slavs; among construction schemes, Europeans, from the linguistic point of
forge-welded combinations of iron and view the youngest and possibly closely
hard steel are represented by about related to Balts. Their ethonogenesis
40%. The Saltovo-Mayatsk culture and geographical craddle is much dis-
dates from the end of the 7th to the cussed - it was possibly the territory be-
early 10th century AD. tween Carpathians and modern central
Sarmatians, Sarmatae in Latin, Sauro- Russia. Ancient Slavs are mentioned in
matoi by Herodotus, an Indo-European Byzantine written sources as Sclavini,
population related to Scythians, al- Antai and Venedoi. During the 6th -
lies of Germanic tribes in wars against 7th centuries Slavs colonized wide re-
Rome (Marcus Aurelius). A Sarma- gions up to the Saale river and Ham-
tian nomadic branch, the Yazygi, pene- burg/Lübeck in Germany before hav-
trated to the Hungarian lowland during ing been pushed back beyond the river
the early Romano-Barbarian period. Nisa by the Germans. During the 8th -
GLOSSARY OF HISTORICAL AND ARCHAEOLOGICAL TERMS 255
9th centuries they began to form their tribes).
own establishments (the realm of Samo, Tacitus, Publius (?) Cornelius, AD ca
then Great Moravia, Kiyv Russia). In 55 - 130. The celebrated Roman his-
the 9th/10th centuries they were split torian. In his De situ, moribus ac pop-
into two entities by the expansion of ulis Germaiae or Germania he collected
Magyars and Bavarians: the southern data on population in the Germania
Slavs (Serbs, Slovenians, Croatians) re- libera which could have been used in
mained to be settled in the Balkans. military campaigns. The work repre-
The common Slavic language survived sents an important source for archaeol-
up to that time before having been dif- ogists. There are some remarks on iron,
ferentiated. as well. However, fundamental works
Sophocles (495 - 406 BC), a renowned by Tacitus, compiled later, are his His-
Greek tragedian whose father Sophillus toriae (14 books) and Annals (Ab ex-
owned a smithy manufacture in which cessu Divi Augusti, 15 books) present-
weapons have been produced by qualifi- ing the history of Rome since Galba up
ed slave swordsmiths. In the Sophocles’ to Nero.
poem Aias the quenching of iron/steel Tchernyakhov culture, Ukraine, 2nd
is mentioned (bafé sidéros). - 4th centuries AD. Named after the
spatha, a double-edged long iron sword tumulus cemetery Tchernyakhiv. In
of Franks in the 5th - 7th centuries fact, a complex of many local groups
AD, the predecessor of which was the between the Carpathians and Dnieper
Celtic sword. It was 70cm - 100cm and Seym rivers were involved. Rural
long, sometimes pattern-welded, suit- settlements with sunken-floored huts,
able both for cavalry and infantry biritual burials, hand- and wheel-made
fight. The spathae had ornamented pottery, metallurgy of iron (slag-pit
hilts of various shapes (base for classifi- furnace fields), Roman imports. Sets
cation) and their scabbards were made of iron artefacts were metallographi-
of wood and/or leather and mounted cally examined showing both simple
with metal. The spatha was a model and more advanced techniques. The
for the medieval long sword. bearers of the culture were possibly eth-
Stradonice, the most renowned Celtic nically heterogeneous (Dacians, Sarma-
oppidum in Bohemia, west of Prague, tians, an even early Slavs might have
upon the Berounka river valley. In 1877 been participitating); however, it seems
a find of 200 gold coins attracted the at- that the leading role played the Ostro-
tention, thenafter rummaged by people goths.
and rich find collections were revealed. temenos, in ancient Greece a sacred en-
Small-scale test excavations before the closure of land consecrated to a Deity.
World War I and II, more intensive It comprised a sacred grove, altar or
field work in the 1980’s. Within the even sanctuary. Sometimes it was to-
acropolis and bailey (up to 90ha) farm- tally unaccessible to the public. In the
steads and wells were found, dwellings Hellenistic period the property of tem-
and workshops producing glass pearls, ple.
enameled ware and metal objects. Ex- Teutons see Germanic tribes.
tended finds of ironwork, mintage at- Theophilus or Schedula diversarum ar-
tested. The final phase of the op- tium, an early medieval handbook de-
pidum coincides with the end of the scribing various techniques of gold- and
pre-Christian era (arrival of Germanic glassworking, making mosaics, bronze
256 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
foundry etc. What concerns ironwork- vicus, a Roman unfortified rural village
ing there are described smith’s tools, (or a town part as well). Many work-
bellows and the technique of carburiz- shops incl. smithies were housed in the
ing and quenching of steel files. The vici which had their own forum. Mar-
date of the work is 11th century AD kets took place there.
and the authorship is, according to Viking Age, in Scandinavia the period
some scholars, ascribed to a West- between about AD 800 - 1200. Vikings
phalian monk Theophilus (Rogerus von were Scandinavian Germanic tribes the
Helmershausen). coastal and inlet settlements (vik) of
Timber grave culture, srubnaya kul- which were bases of these skilled sea-
tura, the latest phase of the Ochre farers, merchants and artisans. The
grave culture in Ukraine and middle raids of Vikings stigmatized a great
Volha river region. Inhumations, at part of the coastal Europe. Some
that time rarely powdered with ochre, Viking groups, as Normans, settled in
were buried in timbered chambers in- the coastal France (Normandy) and
side the burial mounds. The dating England. In the working of iron they
covers ca 1500 to 800 BC. Rare and applied advanced techniques.
small iron objects were revealed in early villa rustica, a Roman private farmstead
phases whilst subsequently the use of of medium size, run by the owner’s fam-
iron (weapons) became more frequent. ily and its slaves tending to a consider-
Ulfberht, a renowned Frankish sword- able autarchy (agriculture, crafts incl.
smith who lived about AD 800 and smiths) and representing an economic
worked in his smithy somewhere on the complex. Just certain luxurious goods
middle Rhine or Maas rivers. His out- were imported from town. Villae rusti-
standing swords were found in Scandi- cae were spread over Roman provinces
navia and Baltic regions and the name as well. From the 2nd century AD
ULFBERHT as a mark of quality was many of them converted to larger lati-
used after his death as well. It was fundiae.
composed, as inlays, of small pattern- votive deposit se sacrified deposit.
welded sticks or etched (as the so-called Weyland the Smith (in German
St. Stephan’s sword in the treasury of Wieland, Nord. Válundr), the leg-
the Prague cathedral). endary master smith, son of the sea
Urartu, also Biainu, Nairi, an ancient giant Wal. The events took place not
realm on the territory of modern Arme- only in the North but also in Frankish
nia, north-eastern Anatolia and north- territories and elsewhere as told in the
ern Iran (9th - 6th centuries BC). It Amelungenlied and in a more abridged
paid taxes in iron to Assyria in the 9th form in the Edda, the collection of
century BC. mythological poems (both formed from
Vendel period. Vendel in Uppland, Swe- the 7th to the 11th century AD). Wey-
den, is a site with pre-Viking richly land learned his art from dwarfs in the
equipped boat graves (precious hel- Mount Glockensachse (believed to be
mets and swords) which gave name to the Caucasus). In a competition with
the time and style. First-class handi- Amilias, the smith of the king Neid-
craft works were documented, together ing, he won and created the famous
with relevant workshops, by the results sword Mimung, having used iron filings
of the excavations at Helgö (see this treated with the nitridic bird-dung and
item). cereal substances. The sword should to
GLOSSARY OF HISTORICAL AND ARCHAEOLOGICAL TERMS 257
have been the best in the world. Chalybes and on iron weaponry of vari-
Winchester, Hampshire, England. A ous tribes in the NE Asia Minor in that
medieval town on the territory of a work.
Celtic site. Evidence of metalworking Zarubincy culture, Ukraine, named af-
- a bell casting place (moulds) serv- ter Zarubincy, Zarubintsy, a cemetery
ing during the construction of the later SE of Kiyv. The spread occuppies the
pulled down cathedral; sets of artefacts territories between the Pritpet, Prut,
from the site were metallographically and middle Dnieper rivers. The burial
investigated showing the application of rite was cremation. The chronology
the technique of welding-together iron is still established vaguely (1st century
and steel. BC to 2nd century AD). Speculations
Xenophon (ca 430 - 355 BC), a Greek about a pre-Slavic or proto-Slavic pop-
historian, author of works dealing with ulation. Villages ran agriculture, in
historical, military, economic and polit- northern parts were some simply for-
ical themes, a pupil of Socrates. In the tified sites. Both free-standing and
context of the history of iron as crucial sunken-floored huts. From the Dnieper
work appears the Anabasis, the home- area sets of iron artefacts were metallo-
march of 10 000 Greek mercenaries hav- graphically investigated. Simple tech-
ing left the service to Persian kings af- niques represent about 80% (iron or
ter severe conflicts in Persia. Xenophon steel).
left valuable information on ironmaking
258 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

ACKNOWLEDGEMENTS

Delivering my contribution devoted to the early history of European blacksmithung I have


to express my thanks to scholars who supported me with moral and actual help. First, to
Henry F. Cleere, the initiator of the project ‘Iron in Archaeology’. The volume dealing
with the ‘European Bloomery Smelters’ I have finished and published, with his support,
in 2000. The second theme, the work of early European blacksmiths, is the subject of the
presented book. Then, my thank belongs to Peter and Susan Crew - both provided me
not only with consulstive help but also with some publications which I have lost during
the flood of 2002; the same concerns Janet Lang from the British Museum.
I am rendering homage to late scholars who, in course of my work, still provided
valuable advices and publications as well, Ronald Frank Tylecote, Robert Thomsen and
Dirk Horstmann.
I owe my thanks to numerous colleagues, technicians, natural historians, historians
and archaeologists who supplied me with data, books, papers and all what belongs to
a fruitful collaboration: Hans-Gert Bachmann, P. Benoit, Darina Bialeková, Kazimierz
Bielenin, Vagn Buchwald, Paul Craddock, Claude Domergue, Gerry McDonnell, Arne
Espelund, Philippe Fluzin, János Gömöri, Andreas Hauptmann, Eva Hjärthner-Holdar,
Ineke Joosten, Ingo Keesmann, Hans-Ludwig Knau, Marc Leroy, Robert Maddin, Irmelin
Martens, L’ubomı́r Mihok, Elżbieta Nosek, Jerzy Piaskowski, Karol Pieta, Vincent Piggot,
Lyudmila Rozanova, Vincent Serneels, Boris Shramko, Gerhard Sperl, Harald Straube,
Nataliya Terekhova, Marco and Constanza Tizzoni, Estanislau Tomàs i Morera, Mercedes
Urteaga, George Varoufakis, Olfert Voss, Allan Williams, Ünsal Yalçin. From my own
country I have to name Jiřı́ Hošek, Jana Siegelová, Věra Souchopová, Vladimır Ustohal,
Natalie Venclová.
In the Archaeological Institute, Prague, assisted me Luboš Jiráň, the Director, my
colleagues Helena Komárková, Dagmar Čerychová, Kateřina Macková-Vytejčková, Hana
Česalová and what concerns computer procedures Magda Mazancová, and Čeněk Čı́šecký.
They deserve my cordial gratitude as well as Stewart Aitchison from Netolice who cor-
rected my English text.
Last but not least I have to give my thanks to my daughter Johana Pleinerová-
Brokešová who took care of the setting an typographical arrangement and to my wife
Ivana not only for understanding but also for virtual work. The work would not be
possible to finish without GAČR grant project No. 404/05/2063.
BIBLIOGRAPHY 259
BIBLIOGRAPHY

The wording of existing English, French and German summaries in items written in
Scandinavian and Slavic languages appears as in the original.
Abels, B.-U. 1989: Ein frühlatènezeitlicher Depotfund von Heidelberg bei Schweinthal,
Gem. Egloffstein, Ldkr. Forchheim, Oberfranken. In: Das archäologische Jahr in Bayern
1988 [l989], 53.

Abrahamsen, N. 1965: Arkaeomagnetism og Jernalderslagge. Kuml 1965, 115 - 131.

Agricola, G. 1556: De re metallica libri XII. Transl. B. Ježek and J. Huml. Praha
1933, reprint 1976. Nat. Tech. Museum, Praha.

Aleksiyev, J. 1976: Nakhodka ot or’diya truda i v’bor’zheniyi ot Carevec. Muzey i


pametnici na kulturata 16/2, 33 - 36.

Allen, D. 1967: Iron Currency Bars in Britain. Proccedings of the Preh. Society
(London) 33, 307 - 325.

Allen, J. R. L. 1986: Interpretation of some Romano-British smithing slag from Awre


in Gloucestershire. HM 20/2, 97 - 104.

Anderson, T. J., Aguston, C., Duvauchelle, A., Serneels, V., Castello, D.


2003: Des artisans à la Campagne. Carrières des meules, forge et voie gallo-romaine à
Chables FR). Fribourg.

Angus, N.S., Brown, G.T., Cleere, H.F. 1962: The iron nails from the legionary
fortress at Inchtuthill, Perthshire. JISI 200, 956 - 968.

Ankner, D. 1996: Die Damaszierung der Spathen aus Altenerdig. In: Das Rei-
hengräberfeld in Oberbayern II. Mainz, 144 - 155.

Anstee, J. W., Biek, L. 1961: A Study in Pattern-Welding. Medieval Archaeology


5, 71 - 93.

Antein, Anteyn see Anteins.

Anteins, A. 1957: Ķentes pilskalna dzels un tērauda izstradajumu strukturas, ipašibas


un izgatavošanas technologija. Archeologija un etnografija I (Riga), 45 - 50.

Anteins, A. 1960a: Dzels un tērauda izstrādājumu strukturās, ipašibas un izgatavošanas


šenaja Latvija. Archeologija un etnografija II, 3 - 60.

Antein, A. K. 1960b: Nakonechniki kopiy iz damasskoy stali drevney Latvii. Tezisi


dokladov. Riga, 39 sq.
260 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Antein, A. K. 1961: Struktura, svoystva i tekhnologiya zheleznykh i stal’nych izdeliy
asotskogo gorodichtcha. AN Lab SSSR. Moskva, 175 sq.
Anteyn, A. 1962a: Nakonechniki kopiy iz svarochnoy uzorchatoy (damasskoy) stali,
naydennyye v Estonii. Eesti NSV teraduste akadeemia trimetisad XI/4. Izvestiya akademii
nauk estonskoy SSR (obchestvennyye nauki), 354 sq.

Anteins, A. 1962b: Senas Latvijas damascata tērauda ṡkēpu gali. Archeologija un


etnografija (Riga) IV, 33 - 46.

Anteyn, A. K. 1963: Nakonechniki kopiy iz svarochnoy uzorchatoy (damasskoy) stali


v drevney Pribaltike. Sov. arkh. (Moskva) 1963/4, 165 - 178.

Anteins, A. 1964: Damascētu un ierakstu zobeni Latvija un to asmeņu. Par technikas


vesturi Latvias PSR (Riga), 65 - 95.

Anteins, A. 1966: Im Ostbaltikum gefundene Schwerter mit damaszierten Klingen.


Waffen- und Kostümkunde (München), 111 - 115.

Anteins, A. K. 1968: Structure and manufacture techniques od pattern-welded ob-


jects in the Baltic States. JISI 206, 563 - 571.

Antonescu, I. 1968: Depozitul de objecte de fier din comuna Negri-Baca̧u si im-


plicaṫiile sale istorice. Le dépôt carpique de Negri et ses implications historiques. Carpica
I (Bacău), 189 - 196.

Arbman, H. 1937: Schweden und das Karolingische Reich. Stockholm.

Archaeometallurgy of Iron 1967-1987. International Symposium of the CPSA Liblice


1987 (R. Pleiner ed.). Prague 1989.

Archaeometallurgy in Europe, vol. 1 and 2, 2003. Proceedings of the International


Conference. Milano.

Archaeometallurgy in The Central Europe III (L’. Mihok and E. Miroššayová eds.).
Acta Metallurgical Slovaca 7/2, Košice 2001.

Archäometallurgie von Kupfer und Eisen in Westeuropa. Symp. Mainz 12 - 15 Sept


1986. Jahrbuch des Römisch-Germanischen Zentralmuseums in Mainz 35 1988 [1991].

Arne, T. J. 1962: Was bedeutet das Vorkommen von Nickel in frühgeschichtlichen


eisernen Gegenständen. Fornvännen 57, 335 - 337.

Arwidson, G. 1977: Valsgärda 7. Die Gräberfunde von Valsgärda III. Uppsala, Lund.

Aspects of Early Metallurgy (W. A. Oddy ed.). London 1977.

Aux origines de la métallurgie du fer en Afrique. Une ancienneté méconnue - Afrique


BIBLIOGRAPHY 261
de l’Ouest et Afrique centrale (H. Bocoum ed.). Edition UNESCO. Paris 2002.

Awty, B. G. 1996: Early cast irons and the impact of fuel availability on their pro-
duction. HM 30/1, 17 - 22.

Åström, p., Maddin, R., Muhly, J. D., Stech, T. 1986: Iron Artifacts from
Swedish Excavations in Cyprus. Opuscula Atheniensia 16/3, 27 - 41.

Baatz, D. 1991: Die schweren Eisenträger von der Saalburg - Zur Form, Funktion und
Metallurgie. With contributions by A. Hauptmann and R. Maddin. Saalburg Jahrbuch
46, 25 - 46.

Bachmann, H.-G. 1970: Eisenschlacken von Eski Kâhta (Vil. Adiyaman), Südtürkei.
Archiv f. d. Eisenhüttenwesen 41, 731 - 736.

Bachmann, H. G. 1982: The identification of slags from archaeological sites. Insti-


tute of Archaeology Occasional Publication No. 6. London.

Bachmann, H. G. 1987: Schlacken: Indikatoren archäometallurgischer Prozesse.


In: Mineralische Rohstoffe als kulturhistorische Informationsquelle (H. W. Hennicke ed.).
Verein Deutscher Emailpalleter. Hagen, 66 - 103.

Backer, G., Dick. W. 1967: Römerzeitliche eiserne Münzprägestempel aus Trier.


Archiv f.d. Eisenhüttenwesen 38, 351 - 356.

Bailly-Mâitre, M.-C. 1987: Brandes en Oisans, Haute Isère: un village minier de


haute montagne au Moyen Age. Huez.

Bailly-Mâitre, M.-C. 1995: Mines et forges au Moyen Age et en début du temps


modernes. Reflection autour de l’example de Brandes. In: Paléometallurgie du fer &
Cultures, 333 - 347.

Bailly-Mâitre, M.-C., Ploquin, A. 1993: Brandes en Oisans. Archéologie et


paléometallurgie d’un village minier au Moyen Age. In: Montanarchäologie in Europa
(H. Steuer and U. Zimmermann eds.). Sigmaringen, 443 - 460.

Balabanov, T. 1981: Zhelezarska i mednikarska robotilnica v Pliska. Muzei i pamet-


nici za kulturata (Sofia) 4, 34 - 39.

Barnard, R., Tamotsu, S. 1975: Metallurgical Remains of Ancient China. Tokyo.

Barner, W. 1935: Ein spätkarolingisches Bauerngehöft auf der Wüstung Assum (Feld-
mark Eime, Kreis Alfeld). Die Kunde III 7/8, 113 - 128 .

Bartošková, A. 1986: Slovanské depoty železných předmětů v Československu. Slaw-


ische Hortfunde von Eisengegnständen in der Tschechoslowakei. Studia Archeol. ústavu
v Brně. Československá akademie věd XIII/2. Praha.
262 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Bartuška, M., Pleiner, R. 1965: Untersuchungen von Baustoffen und Schlacken aus
den frühgeschichtlichen Rennöfen Böhmens und Mährens. In: Technische Beiträge zur
Archäologie 2. Mainz, 1 - 37.

Bartuška, M., Pleiner, R. 1968: Untersuchung der Schlackenproben von Mühlthal.


In: Dannheimer 1968, 97 - 101.

Basler, D. 1977: Rimski metalurški pogon i na naselje v dolini Japre. Römisches


Eisenwerk und Ansiedlung im Japra-Tal. Glasnik zemaljskogo muzeja u Sarajevo, Arch.
N.S. 30/31, 121-215.

Beck, L. 1884: Geschichte des Eisens in technischer und kulturgeschichtlicher Beziehung.


Braunschweig.

Becker, C. J. 1980, 1982: Ein Einzelhof aus der jüngeren vorrömischen Eisenzeit in
Westjütland. Offa 37 (Festschrift Hinz); 39, 53 - 72.

Becker, G. 1961: Niedrig schmelzende Eisen-Arsen-Legierungen für den Verbund


römischer Schwertklingen. Archiv f.d. Eisenhüttenwesen 30, 661 - 665.

Beiträge zum römischen Oberwinterthur - VITUDURUM 7. Ausgrabungen im Unteren


Bühl. Funde aus Metall. Ein Schrank mit Lararium des 3. Jahrhunderts. Zürich, Egg
1996.

Belhoste, J.-F. 1991. L’implantation d’une sidérurgie bergamasque au début du


XVIIe siècle. In: Dal basso fuoco all’altoforno. Sibrium 20 (Brescia), 265 - 274.

Bencard, M. et al. 1979: Wikingerzeitliches Handwerk in Ribe. Acta Archaeol.


Københaven 49 1978 [1979], 113 - 138.

Benoit, P. 1997: La mine de Pampailly XVe - XVIIIe siècles, Brussieu - Rhône.


DARA vol. 14. Lyon.

Benoit, P sine anno: Un site industriel medieval: L’abbaye de Fontenay. Mémoires


de la Commission des Antiquités du département de la Côte-d’Or 36, 219 - 247.

Berg, G. 1955: A Tool Chest in the Viking Age. The ”Mästermyr” find in Gotland.
Årbok Univ. Bergen.

Berciu, D. 1964: Pour une voie Cimmerienne de diffusion de la métsllurgie du fer.


Archeol. rozhledy (Praha) 16, 264 - 279.

Bergmann, K., Billberg, I. 1976: Metallhandwerk. In: Uppgräft förflutet i Lund


(A.W. Mårtensson ed.). Archaeologia Lundensia VII. En investering i arkeologi, Kul-
turhistorisk Museet i Lund.
BIBLIOGRAPHY 263
Bialeková, D. 1981: Dávne slovanské kováčstvo. Old Slav Smithing. Tatran, Bratislava.

Bialeková, D., Mihok, L’., Pribulová, A., Hollý, A., Strečan, V. 1999: Met-
allographic analysis of axe shaped currency bars from Vel’ký Klı́ž and Pobedim. In:
Archaeometallurgy in The Central Europe, 96 - 107.

Biborski, M., Kaczanowski, P., Kȩdzierski Z., Stȩpiński, J. 2002: Badania nad
technologia̧ mieczy z mlodsiego okresu przedrzymskiego i z okresu kultury przeworskiej.
In: Varia Barbarica in honorem Z. Wożniak. Warszawa, Lublin, 81 - 104.

Biborski, M., Kaczanowski, P., Kȩdzierski, Z., Stȩpiński, J. 2003: Manufacur-


ing technology of double and single edged swords from the lst century BC and 2nd century
AD. In: Archaeometallurgy in Europe. Milan, 97 - 107.

Bielenin, K., Mangin, M., Orzechowski, S. 1996: La sidérurgie ancienne et


l’exploitation miniére dans la Montagne Saint-Croix (Petite Pologne) II. Ateliers, habitat,
chronologie. Dialogue d’Histoire Ancienne 22/1, 327 - 373.

Biringuccio, V. 1540: De la Pirotechnia libri X. Transl. M. T. Gaudi and C.S.Smith,


MIT Press, Cambridge, Mass. 1966.

The Birth of the Metallurgy of Iron and the Beginning of the European Iron Age. Acta
PACT 30-3. Strassbourg.

Bjorkman, J. K. 1973: Meteors and Meteorites in the Ancient Near East. Tempe,
Arizona.

Blackburn, T. P. D., Edge, D. A., Williams, A. R., Adams, C. B. T. 2000:


Head protection in England before the First World War. Neurosurgery 47/6, 1261 - 1286.

Blindheim, Ch. 1963: Smedegraven fra Bygland i Morgedal. Viking 26, 25 - 80.

Bloomery Ironmaking during 2000 years I-III. Seminar in Budalen. Trondheim, 1991 -
1993.

Blümner, H. 1886: Technologie und Terminologie der Gewerbe und Künste bei
Griechen und Römern IV. Leipzig.

Böhne, C. 1963: Die Technik der damaszierten Schwertern. Archiv f. d. Eisenhüttenwesen


34, 227 - 234.

Böhne, C., Dannheimer, H. 1961: Studien an wurmbunten Klingen des frühen


Mittelalters. Bayerische Vorgeschichtsblätter 26, 107sq.

Boll, P.O., Erismann, T.H., Muster, W.J. 1981: Metallkundliche Untersuchung


eines frühen mitteleuropäischen Eisenschwertes. In: Frühes Eisen in Europa, 45 - 51.
264 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Boll, O., Flisch, A., Senn-Luder, M. 1999: Archäeometallurgische Untersuchun-
gen an keltischen Waffen von Wartau Ochsenberg (SG). In: Universitätsforschungen zur
prähistorischen Archäologie Vol. 55 (T. Della Cosa ed.). Bonn.

Borkovský, I. 1957: K výkladu nožů na slovanských pohřebištı́ch. Zur Interpretation


der Messer in den slawischen Gräberfeldern. Archeol. rozhledy 9, 543 - 560.

Boroffka, N. 1991: Die Verwendung von Eisen in Rumänien von den Anfängen bis
zum 8. Jahrhundert v. Chr. Berlin.

Böttger, B. 1975: Die Landwirtschaft. In: Die Römer an Rhein und Donau (K.
Günther and H. Köpstein eds.). Berlin, 138 - 188.

Bouchayer, J. 1956: Les Chartreux mâitres des forges. In: Le fer à travers les âges.
Nancy, 143 - 159.

Bouzek, J. 1978 Zu den Anfängen der Eisenzeit in Mitteleuropa. Zeitschrift f. Archäologie


(Berlin) 12, 8 - 14.

Brepohl, E. 1987: Theophilus Presbyter und die mittelalterliche Goldschmiedekunst.


Leipzig.

Brewer, C. W. 1976: Metallographic examination of six ancient steel weapons. HM


10/1, 1 - 10.

Brewer, C. W. 1981: Metallographic examination of Medieval and Post-Medieval


Iron Armours. HM 15/1, 1 - 8.

Brown, P. D. C. 1971: A hoard of currency bars from Appleford, Berks. Proceedings


of the Preh. Society 37/1, 226 - 228.

Brunaux, J. L., Mentel, P., Rapin, A. 1980: Une sanctuaire gauloise à Gournay-
sur-Aronde (Oise). Gallia 38/1, 1 - 25.

Brunnengräber, C., Gaitzsch, W. 1993: Antikes Eisen-Konservierung und Auswer-


tung eines umfangreichen römischen Hortfundes aus Aldenhoven Pattern, Kr. Düren. Das
Rheinische Landesmuseum Bonn 3, 59 - 62.

Bubenı́k, J. 1972: K problematice železné misky tzv. slezského typu. Zur prob-
lematik der Eisenschale vom sog. Schlesischen Typus. Archeol. rozhledy (Praha) 24, 542
- 567.

Buchner, G. 1969: Mostra degli scavi di Pithecusa. Dialogi di Archeologia 1-2, 85 -


101.

Buchwald, V. F. 1993 [1994]: Smedejern, Essesvejsning og slaggekarakterisering.


Danish Metallurgical Society, November 1993 [January 1994].
BIBLIOGRAPHY 265

Buchwald, V. F. 1999: Blaesterjern, Kloder og Klimpjern - smedejern fremstillet ved


direct processer i blaesterovne. In: Klimp og Kloder. Jern i middelalderens Danmark (P.
H. Jensen ed.), Thorning-Kjellerup.

Buchwald, W. F. 2001: Ancient Iron and Slags in Greenland. Meddelser om Grønland.


Man and Society vol. 26, Copenhagen.

Buchwald, W. F. 2002: Jern og stål i Skandinavien før 1600 belyst ved slaggenana-
lysenmetoden. Dansk Metallurgisk Selskab, Kolding Fjord.

Buffat, L., Petitot, H. 1998: Une activité métallurgique tardo-antique su l’établissement


de Mayran (Saint-Victor-La-Coste, Gard). In: Recherches sur l’économie du fer. 175 -
180.

Bühler, H.-E., Strassburger, Chr. 1966a: Werkstoffkundliche Untersuchungen an


zwei fränkischen Schwertern aus dem 9. Jahrhundert. Archiv f.d. Eisenhüttenwesen 37/8,
612 - 619.

Bukowski, Z. 1981a: Die ältesten Eisenfunde und älteste Eisengewinnung im Bereich


der Lausitzer Kultur im Flussgebiet von Oder und Weichsel. In: Frühes Eisen in Europa,
69 - 77.

Bukowski, Z. 1981b: Najstarsze znaleziska przedmiotów żelaznych w środkowej Eu-


ropie a pocza̧tki metalurgii żelaza v kulturze lużyckiej w dorzeczu Odry i Wisly. Die
ältesten Eisenfunde in Mitteleuropa und die Anfänge der Eisenmetallurgie in der Lausitzer
Kultur im Stromgebiet der Oder und Weichsel. Archeologia Polski 26/2, 321 - 401.

Bulard, A. 1978: Découverte exceptionelle - un dépôt d’armes du 2e âge du fer.


Archéologia No 115, 70 - 71.

Bulliot, J. G. 1899: Fouilles du Mont Beuvray (ancienne Bibracte) de 1867 à 1895.


Autun.

Burov, V. A., Rozanova, L. S. 2003: K tekhnologicheskoy kharakteristike kuznech-


nykh izdeliy s selichtcha kul’tury dlinnykh kurganov Gorodsk-1 na reke Shlime (Valday).
Technological characteristics of blacksmith’s production from the settlement of long bar-
rows culture Gorodsk-1 on the Shlime river. Kratkie soobchtcheniya IA RAN 214, 104 -
111.

Burton Brown, T. 1950: Iron objects from Azerbaijan (analyses by A. Herbert).


Man 50, 1 - 9.

Cabalska, M. 1964: Wyniki dotychczasowych badań archeologicznych w Maszkow-


icach, pow. Nowy Sa̧cz. Wiadomości Archeol. 30/1-2, 121 - 128.

Campbell, J., Fahy, F. 1984: A metallurgical study of iron pile shoes from the Ro-
266 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
man bridge at Minturnae. HM 18/1, 21 - 30.

Čangova, J. 1962: Srednevekove or’diye na truda v B’lgaria. Outils du Moyen Age


en Bulgaria. Izvestiya na Archeologicheski institut (Sofia) 25, 19 - 55.

Capelle, T. 1980: Bemerkungem zum isländischen Handwerk in der Wikingerzeit und


im Mittelalter. Frühmittelalterliche Studien 14, 423 - 436.

Capelle, T. 1994: Die Miniaturkette von Szilagysomlyó (Ṡimleul Silvanici). Univer-


sitätsforschungen zur prähistorischen Archäologie (Bonn) vo. 22.

Carpenter, H. C. H., Robertson, J. M. 1930: The Metallography of some Ancient


Egyptian Implements. JISI 121, 417 - 454.

Cech, B. 2000: Gold and silver production in the fifteenth and sixteenth century based
on results of archeological excavation in the Gastein Tal, Austria. In: Iron, Blacksmiths
and Tools, 20 - 33.

Cech, B., Wallach, G. 2000: Interdisciplinary research on a miner’s smithy of the


16th century in Gastein, Salzburg, Austria. In: Il ferro nelle Alpi, 114 - 123.

Čech, P. 1998: Early medieval smithies at Žatec, NW Bohemia. Archeol. rozhledy


50, 280.

Chakrabarti, D. K. 1977: Distribution of iron ores and arcaeological significance of


early iron in India. JESHO 20/2, 166 - 184.

Chambers Science and Technology Dictionary (P. B. M. Walker ed.). Cambridge, Ed-
inburgh, New York, New Rochelle, Melbourne, Sydney 1988.

Chardron-Picault, P., Pernot, M. 1999: Un quartier antique d’artisanat métallurgique


à Autun. Le site du Lycée militaire. dAf vol. 76. Paris.

Charles, J. A. 1980: The Coming of Copper and Copper-Base Alloys and Iron. A
Metallurgical sequence. In: The Coming of the Age of Iron, 151 - 181.

Charles, J. A. 1984: A Middle Bronze Age iron punch of southeast Drenthe. Palaeo-
historia 26, 95 - 99.

Cheng Shih-Po 1958: An Iron and Steel Work of 2000 years ago. China Reconstructs,
January 1958, 32 - 34.

Chidioşan, N. 1980: Depozitul de unelte de fier descoperite ı̂n aş ezacea dacică de la
Tăşad, commune Drageşti, judetul Bihor. Le dépôt d’objets en fer découvert dans le habi-
tat dacique de Tăşad, commune de Drăgeṡti, district de Bihor. Crisia (Oradea) 10, 55 - 64.
BIBLIOGRAPHY 267
Ciglemečki, S. 1991: Poznorimski depo z Rudne pri Rudnici. Der spätrömische Hort-
fund aus Rudna bei Rudnica. Arheološki vestnik 42, 255 - 232.

Čilinská, Z. 1984: Depoty železnych predmetov z konca 8. stor. na Slovensku. Hort-


funde von Eisengenständen aus dem 8. Jahrhundert in der Slowakei. In: Zbornı́k prác
L’. Kraskovskej (k životnému jubileu). Bratislava, 163 - 172.

Cirkin, A. V. 1971: Nizhneborkovskiy klad. Sov. arkh. 1971/3, 276 - 281.

The Civilisation of Iron. From prehistory to the third millennium (W. Nicodemi, ed.).
Olivares, Milano 2004.

Čižmář, M. 1990: Časně laténské nálezy z hradiska ”Černov”, okres Ježkovice.


Frühlatènezeitliche Funde aus dem Burgwall ”Černov”, Gemeinde Ježkovice. In: Pravěké
a slovanské osı́dlenı́ Moravy. Sbornı́k k 80. narozeninám Josefa Poulı́ka. Brno, 196 - 204.

Čižmář, M. 1993: Frühlatènezeitlicher Burgwall ”Černov” in Mähren (Tschechische


Republik). Arch. Korrespondenzblatt 23, 267 - 282.

Cleere, H. F. 1958: Roman Domestic Ironwork as illustrated by the Brading, Isle of


Wight, Villa. Bull. of the Inst. of Archaeology 1, 55 - 74.

Coghlan, H. H.: 1956: Notes on the Prehistoric and Early Iron in the Old World.
Oxford. 2nd edition 1977.

The Coming of the Age of Iron (T. A. Wertime and J. D. Muhly eds), Yale University
Press. New Haven, London, New York 1980.

Comşa, M., Constantinescu, Gh. 1969: Depozitul de unelte şi arme din epoca
feudală timpurie descoperit la Dragoslaveni (jud. Vrancea). Dépôt d’outils et d’armes de
la haute époque féodale découvert à Dragoslaveni, département de Vrancea. SCIV 20,
425 - 436.

Comşa, M., Gheannopoulos, H. 1969: Unelte şi arme din epoca feudală timpurie
descoperte la Radovanu (jud. Ilfov). Outils et armes en fer (accompagmés de tessons
céramiques) sur le territoire de la commune de Radovanu (Roumanie). SCIV 20/4.619 -
621.

Conin, A., Guillot, I., Grandemagne, J., Benoit, P. 1995: La forge du Samson.
In: Paléometallurgie & Cultures, 375 - 386.

Craddock, P. 1992: Analysis of the two objects from the Ross expedition of 1818. In:
Aspects of Early North American Metallurgy (M. L. Weyman, J. C. H. King, P. Craddock
eds). British Museum Occasional Paper 79. London.

The Craft of the Blacksmiths (H. Cleere, B. G. Scott eds). Belfast 1987.
268 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Crew, P. 1991: The Experimental Production of Prehistoric Bar Iron. HM 25/1, 21
- 36.

Crew, P. 1992: Aspects of the iron supply. Study 13. In: Danebury: an Iron Age
hillfort in Hampshire 6. CBA Research Report 102, 276 - 284.

Crew, P. 1994: Currency Bars in Great Britain. Typology and Function. In: La
sidérurgie de l’Est de la France, 345 - 350.

Crew, P. 1996: Bloom refining and smithing slags and other residues. Historical Met-
allurgy Society: Archaeology Datasheet No 6.

Crew, P. 2000a: The influece of clay and charcoal ash on bloomery slags: In: Il ferro
nelle Alpi Iron in the Alps, 38 - 48.

Crew, P. 2000b: Iron objects. In: Hougton Down, Stockbridge, Hants, 1994 (B. Cun-
liffe, C. Plas eds) Oxford Monograph 49, 107 - 112.

Crew, P., Salter, Chr. 1993: Currency bars with welded tips. In: Bloomery Iron-
making During 2000 Years III. Trondheim, 11 - 25.

Crişan, I. H. 1960: Un depozit de unelte descoperit la Lechinţa de Mureş (plugul la


geto-daci). Dépôt d’outils agricoles découvert à Lechinta de Mures (la charrue chez les
Gèto-Daces). SCIV 11/2, 285 - 301.

Cunliffe, B. 1972: The Late Iron Age Metalwork from Bulbery, Dorset. The Anti-
quaries Journal 52/II, 293.

Cunliffe, B. 1983: Danebury - Anatomy of an Iron Age Hillfort. London.

Cunliffe, B., Poole, C. 2000: Hougton Down, Stockbridge, Oxford.

Cüppers, J. 1967: Gallo-römischer Bauernhof bei Horath, Kr. Bernkastell. Trierer


Zeitschrift 30, 114 - 142.

Curle, J. 1911: A Roman Frontier Post and Its People. The Fort of Newstead in the
Parish of Melrose. Glasgow.

Curta, F. 1997: Blacksmith, Warrior, and Tournament of Value: Dating and In-
terpreting Early Medieval Hoards of Iron Implements. In: Eastern Europe. Ephemeris
Napocensis (Cluj) 7, 211 - 268.

Curtis, J. V., Wheeler, T. S., Muhly, J. ED., Maddin, R. 1979: Neo-Assyrian


Ironworking Technology. PAPS 123/6, 369 - 390.

Czysz, W. 1991: Der Eisendepotfund aus dem römischen Gutshof von Oberndorf a.
Lech. Das archäol. Jahr in Bayern 1990. Stuttgart 1991, 120 - 126.
BIBLIOGRAPHY 269

Dabas, M., Dieudonné-Glad, N., Poirier, P. 2002: Caractérisation des structures


d’une forge antique: Approche archéologique, géophysique et anthracologique. Revue
d’archéometrie 26, 141 - 154.

Da̧browcy, I. and K. 1967: Cmentarzysko z okresów póżnolateńskiego i wplywów


rzymskich w Wesólkach, pow. Kalisz. Cemetery of the Late La Tène and Roman influ-
ences period at Wesólki, Kalisz District. Warszawa, Wróclaw, Kraków.

Daicoviciu, H. et al. 1953: Şantierul Gradiştea Muncelului. SCIV 4/1-2, 153 - 219.

Daicoviciu, H. 1972: Dacia de la Burebista la cucerirea Romană. Dakien von Bure-


bista bis zur römischen Eroberung. Cluj.

Dannheimer, H. 1968: Epolding-Mühlthal. Siedlung, Friedhof und Kirche des frühen


Mittelalters. München.

Davies, O. 1926/27: Report on Excavations at the Toumba and Tables of Vardaróvtsa.


The Annual of the British School at Athens 27, 197sq (analyses by W. I. Cattle, 233).

Delpino, F. 1988: Prime testimonianze dell’uso del ferro in Italia. In: The First Iron
in the Mediterranean, 47 - 68.

Dickmann, H. 1927: Untersuchungen an eisernen Schmiedestücken aus der gallisch-


romanischen Periode. In: L. Frémont: La qualité du fer et les procédés de forgeage de
l’époque gallo-romaine. Génie Civil 90, 198 - 199.

Dickmann, H. 1962: Neue Beiträge zur Geschichte des Drahtziehens. Stahl u. Eisen
82, 166 - 169.

Dieudonné-Glad, N. 1997: Les Oulches - a late Roman iron smelting and smithing
workshop. In: Early Ironmaking in Europe. Plas Tan y Bwlch, Wales, 17 - 20.

Dillmann, P., Bernard, P., Fluzin, P. 2003: Use of iron for the building of me-
dieval monuments. The Palais des Papes in Avignon and other French buildings. In:
Archaeometallurgy in Europe, Milan, 199 - 208.

Dolenz, H. 1998: Eisenfunde aus der Stadt auf dem Magdalensberg. Klagenfurt.

Doncheva-Petkova, L. 1980: Za metalodobiva i metaloobrabotvaneta v Pliska. Arche-


ologiya (Sofia) 22/4, 27 - 36.

Dostál, B. 1978: Zemnice s depotem pod valem hradiska Břeclav-Pohansko. Gruben-


wohnung mit Hortfund unter dem Wall der Wallburg Břeclav-Pohansko. Sbornı́k pracı́
Fil. fak. brněnské univerzity E 22-23, 103 - 134.

Dostál, B. 1983: Železné sekerovité hřivny z Břeclavi-Pohanska. Sbornı́k pracı́ Fil.


270 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
fak. brněnské univerzity E 27. 135 - 198.

Doswald, C. 1994a: Les lingots de fer protohistoriques en Europe occidentale: problématique


générale. In: La sidérurgie ancienne dans l’Est de la France, 333 - 343.

Doswald, C. 1994b: Zum Handwerk der vici. Pro Vindonissa 1993 (1994), 3 - 18.

Drda, P. 2000: Les arts du feu sur les oppida celtiques IIe - Ie s. av. J.-C. In: Les
Celtes et les arts du feu. Métaux, Céramique, Verre. Dossier d’Archéologie No 258, Nov.
2000, 18 - 23.

Drda, P., Rybová, A. 2001. Model vývoje velmožského dvorce 2. - 1. stoletı́ před
Kristem. Modell der Entwicklung des Herrengehöftes im 2. - 1. Jahrhundert v. Chr.
Památky archeol. (Praha) 92, 284 - 349.

Dunikowski, C., Leroy, M., Merluzzo, P., Ploquin, A. 1998: Les déchets
paléometallurgiques: quels indices pour une forge? In: Les métaux antiques, 145 - 152.

Durdı́k, T. 1990: A medieval smithy in the castle of Jindřichův Hradec. Archeol.


rozhledy (Praha) 402 709.

Durman, A. 2002: Iron resources and production for the Roman frontier in Pannonia,
HM 36/1, 24 - 32.

Egan, G. 1996: Some Archaeological Evidence for Metalworking in London c. 1050


AD - c. 1700 AD. HM 30/2, 83 - 94.

Egger, R. 1961: Die Stadt auf dem Magdalensberg - ein Grosshandelplatz. Die
ältesten Aufzeichnungen über den Metallwarenhandel auf dem Boden Österreichs. Graz,
Wien, Köln.

Ehrenreich, R. M. 1987: A Study of Iron Technology in Wessex Iron Age. In: The
Craft of the Blacksmiths, 105 - 112.

Eilender, W. 1933: Untersuchungsbericht über Eisensachen aus dem alamanischen


Friedhof Herten. Badische Fundberichte 3/1, 27 - 31.

Ergebnisse der Ausgrabungen in Manching 1984 - 1987 (F. Meier et al.). Steinau,
Stuttgart 1992.

El Sayed El Gayar 1995: Pre-Dynastic Iron beads from Gerzeh, Egypt. JAMS - In-
stitute of Archaeometallurgy. Studies of the Institute of Archaeology, University College,
London 19/1, 11 - 12.

Eschenlohr, L. 1998: Les ateliers de forgeron de Develier-Courtételle (Jura, Suisse).


In: Les métaux antiques, 19 - 22.
BIBLIOGRAPHY 271
Eschenlohr, L. 2004: Recherches archéologiques sur le district sidérurgique du Jura
central, Suisse. Lausanne.

Eschenlohr, L., Friedli, W., Senn-Luder, M. 1999: Develier-Courtételle (Jura).


Une activité préindustrielle: le travail du fer. Helvetia Archaeologica 119/119, 73 - 82.

Espelund, A. 1992: The Mellager site in Trondheim - a complex of metal workshops


and its role in the Medieval iron metallurgy. In: Bloomery Ironmaking During 2000 Years
II, Trondheim, 93 - 114.

Esperandieu, E. 1910 - 1938: Récueil général des bas-reliefs, statues and bustes de
la Gaule romaine, Paris (volumes I - XI).

Europäische Technik im Mittelalter 820 bis 1400. Tradition und Innovation, ein Hand-
buch (U. Lindgren ed.). Berlin 1997.

Excavations at Helgö V: 1. Workshops, Part II. Stockholm 1978.

Experimental Iron Production - Archaeology, Technology and Experiments (L. Chr.


Nørbach ed.). Historical-Archeological Experimental Centre, Tech. Report No 3. Lejre
1997.

Fedorov, G. V. 1953: Gorodichtche Ekimaucy. KS (Moskva) 50, 104 - 124.

Feldinger, E. M. 1990: Eine gesunkene Schiffsladung des 15./16. Jahrhunderts aus


der Salzach. Archäologie beiderseits der Salzach. Salzburg, 149 - 150.

Fell, V.1993: Examination of four Iron Age ferrous hammer heads from Bredon Hill
(Hereford and Worcester), England. HM 27/2, 60.

Fell. V. 1995: Metallographic examination of Iron Age tools from Somerset. HM


29/1, 1 - 11.

Fell, V. 1997: Iron Age Iron Files from England. Oxford Journal of Archaeology 16/1,
79 - 98.

Fell, V. 1998: Iron Age ferrous hammerheads from Britain. Oxford Journal of Ar-
chaeology 17/2, 207 - 222.

Fell, V. 2003: Metallographic Examination of Two Iron Age Hooked Blocks from
England. In: Prehistoric and Medieval Direct Iron Smelting in Scandinavia and Europe.
Aarhus, 129 - 133.

Fell, V., Salter, Chr. J. 1998: Metallographic examination of seven ferrous axe-
heads from England. HM 32/1, 1 - 6.

Feugère, M., Serneels, V, 1998: Production, commerce et utilisation du fer entre


272 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
l’Ebre et Rhône: Premiers éléments de reflection. In: Recherches sur l’économie du fer,
251 - 263.

Fichtl, S. 2000: La ville celtique (Les oppida de 150 av. J.-C. à 15 ap. J.-C.). Paris.

Filip, J. 1966, 1969, 1998: Enzyklopädisches Handbuch zur Ur- und Frühgeschichte
Europas 1, 2, III (Addenda, J. Hrala ed.), Praha.

The First Iron in the Mediterranean. Proceedings of the Populonia/Piombino Sympo-


sium 1988 (G. Sperl ed.), PACT 21 1988. Strasbourg.

Fischer, F. 1959: Der spätlatènezeitliche Depot-Fund von Kappel (Kreis Saulgau).


Urkunden zur Ur- und Frühgeschichte aus Südwürttemberg-Hohenzollern, Heft 1. Stuttgart.

Fischer, Th., Schmotz, K. 1991: Zwei Metallsammelfunde aus dem Lagerdorf des
Kastells Künzig - Quintana, In: Das archäologische Jahr in Bayern 1990. Stuttgart 1991,
126 - 130.

Flachenecker, H. 1997: Handwerkliche Lehre und Artes mechanicae. In: Europäische


Technik im Mittelalter 800 bis 1400, 493 - 502.

Flinders-Petrie, W. M. 1928: Gerar, London.

Fluzin, P. 1999: Ponte di Val Gabbia III: I reperti metallici della forgia. Primi resul-
tati delle studio metallographico. In: La miniera perduta. Bienno 189 - 194.

Fluzin, P. 2000: Ponte di Val Gabbia III (Bienno). Les premiers résultats des études
métallpgraphiques. In: Il ferro nelle Alpi - Iron in the Alps. Bienno, 24 31.

Fluzin, P. 2002: La chaı̂ne opératoire en sidérurgie: matériaux archéologies et procédés.


Apport des études métallographiques. In: Aux origines de la métallurgie du fer en Afrique,
59 - 91.

Fluzin, P., Leclère, D. 1998: Etat de l’interprétation des scories à partie d’investigations
métallographiques. In: Les métaux antiques, 135 - 144.

Fomin, L. D. 1974: Tekhnika obrabotki zaliza v Ol’vii i Tiri. Arkheolohiya (Kyiv)


12, 25 - 31.

Forbes, R. J. 1950: Metallurgy in Antiquity. Leiden.

Fox, C. 1946: A Find of the Early Iron Age from Llyn Cerrig Bach, Anglesey. Cardiff.

France-Lanord, A. 1949: La fabrication des épées damassées aux époque mérovingienne


et carolingienne. Les Pays gaumais 10, 19 - 45.

France-Lanord, A. 1963: Les lingots de fer protohistoriques. Revue d’histoire de la


BIBLIOGRAPHY 273
sidérurgie IV/3, 166 - 178.

France-Lanord, A. 1964: La fabrication des épées de fer gauloises. Revue d’histoire


de la sidérurgie V/4, 315 - 327.

Francovich, R. 1993: Mining and metallurgical activity in the Campiglia Marit-


tima (Tuscany) and the archaeological excavation at Rocca San Silvestro. In: Montan-
archäologie in Europa, 429 - 442.

Fridriksson, T. Á., Hermanns-Audardótir, M. 1992. Ironmaking in Iceland. In:


Bloomery Ironmaking During 2000 Years II. Trondheim, 5 - 15.

From Bloom to Knife. Internat. Symp. CPSA Kielce-Ameliówka 1989. Materialy


Archeol. (Kraków) 26, 5 - 127.

Frühe Eisengewinnung in Joldelund, Kr. Nordfriesland 2(A. Haffner, H. Jöns, J. Re-


ichstein eds.). Bonn 2000.

Frühes Eisen in Europa. Festschrift W. U. Guyan (H. Haeffner, R. Pleiner eds). Acta
des 3. Symposium des CPSA, Schaffhausen u. Zürich 1979. Schaffhausen 1981.

Fry 1932: Untersuchung eines keltischen Spitzbarrens aus dem Fund von Sauggart
(Kr. Ehringen). Fundberichte aus Schwaben N.F. 12 1930/31, 45 - 50.

Fulford. M., Sim, O., Doig, A., Painter, J. 2004a: Interpretation of single-edged
microstructures in Roman armour metallography. In: Metallography 2004, 641 - 644.

Fulford, M., Sim, D., Doig, A., Painter, J. 2004b: Metallography of Roman
Armour from Northwest Europe. In: Metallography 2004, 83 - 90.

Furmánek, V. 1988: Eisen während der Bronzezeit in der Slowakei. Zeitschrift f.


Archäologie 23, 183 - 189.

Furmánek, V., Veliačik, L., Vladár, J. 1999: Die Bronzezeit im slowakischen


Raum. Rahden/Westf.

Gaál, A. 1982: Csázárkor vasdepot Dombóvár kotárábol. Imperial period iron hoard
from the surrounding of Dombóvár. Communications Archaeologiae Hungaricae (Bu-
dapest) 1982, 73 - 91.

Gaitzsch, W. 1978: Römische Werkzeuge. Kleine Schriften zum Kenntnis der römischen
Besetzung Südwstdeutschlands. Stuttgart.

Gaitzsch, W. 1980: Eiserne römische Werkzeuge. BAR Int. Ser. 71/1 - 2.

Galliou, P. 1984: Iron in the Iron Age and Roman Armorica. In: Offa 48 (Festschrif.
H. Hingst), 77 - 63.
274 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Galuška, L. 1992: Dvě velkomoravské kovárny s depoty ze Starého Města. Two Great
Moravian forging shops with hoards from Staré Město. Časopis Moravského muzea. Acta
Musei Moraviae Sc. Soc. 77, 123 - 160.

Gansiniec, Z. 1956: Żelazny penia̧dz Spartan i geneza obolusa. The iron money of
Spartans and the origin of the Obolus Currency. Archeologia (Warszawa, Wróclaw) 8,
367 - 413.

Ganzelewski, M. 2000: Archäometallurgische Untersuchungen zur frühen Verhüttung


von Raseneisenerzen am Kammberg bei Joldelund, Kreis Nordfriesland. In: Frühe Eisen-
gewinnung in Joldelund 2, 3 - 100.

Garbsch, J. 1966: Der Moosberg bei Murnau. München.

Gardi, R. 1954: Der schwarze Hephästus. Bern.

Garnier, M. 1975: A la recherche du passé antique de Chailly en Brie-Calagum. Bull.


Soc. arch. champenois 68/4, 50 - 57.

Gaspari, A. 1999: Römische Schmiedewerkstätten auf der Hügel Ulaka in Innerkrain,


Slowenien. Die Ausgrabungen von Walter Schmid (1936 - 1940). In: ‘...und sie formten
das Eisen’. Archaeologia Austriaca 82/83 1998/1999, 519 - 523.

Gebhard, R. 1991: Aus der Werkstatt eines antiken Feinschmiedes - Zum Depot von
Ošaniči bei Stolec in Jugoslawien. ZAK 48/1, 2 - 10.

Geiger, T, 1996: Untersuchung von Schlackenfunden aus dem Unteren Bühl (contri-
bution by V. Serneels). In: Beiträge zum römischen Oberwinterthur - Vitudurum 7. Die
Funde aus Metal. Ein Schrank mit Lararium des 3. Jahrhunderts. Zürich, Egg, 210 -
221.

Geisler, H. 1976: Ein Gerätdepot der späten römischen Kaiserzeit aus Breslach, Kr.
Eisenhüttenstadt. Veröffentlichungen de Museums f. Ur- u. Frühgesch. Potsdam 10, 141
- 158.

Gettens, R. J., Clarke, R. J., Chase, W. F. 1971: Two Early Chinese Bronze
Weapons wirh Meteoritic Iron Blades. Freer Gallery of Art, Washington, D. C. Occasional
Papers 4/1.

Gierow, P. G. 1966: The Iron Age Culture of Latium I. Lund, 353 - 359.

Ginalski, J. 1997: Wczesnośredniowieczne depozyty przedmiotów żelaznych z grodziska


”Fajka” w Trepczy kolo Sanoka. Sprawozdania Archeol. 49, 221 - 241.

Ginouvez, O., Pomarèdes, H., Feugère, M. 1998: Le travail du fer sur la villa de
la Domergue à Sauvian (Hérault). In: Recherches sur l’économie du fer, 181 - 185.
BIBLIOGRAPHY 275

Ginouvez, O., Vidal, L. 1998: La forge antique de Saint-Jean d’ Aureilhan à Béziers


(Hérault). In: Recherches sur la l’économie du fer, 150 - 152.

Giot, P. R. 1964: Les lingots de fer bipyramidal de Bretagne. Ann. Bretagne 7, sine
pag.

Glodariu, I. 1980: Ateliers métallurgiques de Sarmizegetuse dace. In: Acta du II


Congrès international de thracologie II. Bucaresti, 82 - 92.

Glodariu, I., Cı̂mpeanu, M. 1966: Depozitul de unelte agricole de la Dedvad. SCIV


17/1, 200 - 231.

Glodariu, I., Zrinyi, A., Gyulai, P. 1970: Le dépot d’outils romains de Mărculeni.
Dacia N.S. 14, 207 - 231.

Glowacki, Z. 1961: Uwagi na temat technologii wykonawczej zabytków metalowych


z XI - XIII w z Ostrówa Tumskiego w Poznaniu (1953 - 1954). In: Poznań we wczesnym
średniowieczu, vol. III. Wroclaw, Warszawa, 95 - 105.

Glowacki, Z. Losiński, W. 1960: Badania metaloznawcze noży w wczesnośrednio-


wiecznego cmentarzyska w Mlodzikowie, pow. Środa. Fontes Archaeol. Posnanienses 11,
166 - 178.

Goldenberg, G. 1993: Frühe Blei-, Silber- und Kupfergewinnung im Südschwarzwald.


Hüttenplätze und Bergschmieden. In: Montanarchäologie in Europa, 231 - 248.

Goldenberg, G. 1999: Die Erzlagerstätten im Sulzburger Tal. Arch. Nachrichten


aus Baden, Heft 61/62, 13 - 22.

Golschani, F., Hellermann, B. A., Keesmann, I. 1988 [1991]: Schlacken und


Eisenverarbeitung von Toscanos - Morro de Mezquitilla, Provinz Málaga, Südspanien. In:
Archäometallurgie von Kupfer und Eisen in Westeuropa, 550 - 552.

Gomolka, G. 1975: Die Metallproduktion. In: Die Römer am Rhein und Donau.
Berlin, 189 - 230.

Gömöri, J. 1981: A korai középkori vasolvasztó kemencék és az ékelt vasbucák


kérdése. On the problem of early medieval iron smelting furnaces and split iron pigs.
In: Iparégészet - Industrial Archaeology I (Sopron), 109 - 121.

Gömöri, J. 1995: Archäologische Überreste alter Schmiedewerkstätten in Ungarn. In:


Paléometallurgie de fer & Cultures, 187 - 201.

Gömöri, J. 2000: Az avar kori és Árpád kori Vaskohászat régészeti emlékei Pan-
noniában. The archaeometallurgical sites in Pannonia in the Avar- and early Árpád
period. Sopron.
276 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Gordon, R. B., van der Merwe, N. J. 1984: Metallographic Study of Iron Arte-
facts from the Eastern Transvaal. Archaeometry 24/1, 108 - 127.

Götze, A. 1928: Die ”Schwurschwerter” der Wartburg - Taleae ferreae. Mannus


(Leipzig), Ergänzungsband 6, 138 - 144.

Grakov, V. N. 1977: Ranniy zheleznyy vek (kul’tury zapadnoy i jugo-vostochnoy


Evropy). The early Iron Age (Cultures of Western and South-Eastern Europe). Moskva.

Gralfs, B. 1988: Metallverarbeitende Werkstätten in Pompeji. BAR Int. Ser. 433.


Oxford.

Grandemagne, J. 1990: Le Samson, atelier et habitat d’une mine d’argent au XVIe


siècle. Pierres et Terre 1990, 116 - 120.

Gregory, T. 1976: A hoard of late Roman metalwork from Weeting, Norfolk. Norfolk
Archaeology 36, 263 - 272.

Grimm, P. 1990: Tilleda - eine Königspfalz am Kyffhäuser 2. Die Vorburg u. Zusam-


menfassung. Berlin.

Gruat, P. 2001: Scène de forge. L’Archéologue - Archéologie nouvelle No 50 (Août -


September), 30.

Guillot, I., Bertin, F., Fluzin, P., Béranger, G. 1995: La pointerolle, outil des
mineurs du XVIe siècle. Synthèse des études métallographiques. In: Paléometallurgie &
Cultures, 509 - 524.

Guillot, I., Fluzin, P. 1987a: Interprétation structurale d’élaboration et de l’utilisation


d’outils miniers. Bulletin de la Socété Préhistorique Française 84, 248 - 256.

Guillot, I., Fluzin, P., Clavel, M., Béranger, M. 1987b: Structure des outils
miniers du XVe et du XVIIIe siècle. Interpretation thermomécanique. Matériaux et
Techniques - Octobre 1987, 411 - 419.

Gurin, M. F. 1982: Drevneye zhelezo beloruskogo Podneprovya. Minsk.

Gurin, M. F. 1984: Issledovaniye trechpolosnykh nozhey polockoy zemli. Unter-


suchung von Dreischichtmessern aus dem Polock-Gebiet. Slovenská archeológia 32/2, 311
- 326.

Gurin, M. F. 1987: Kuznechoye remeslo Polockoy zemli (IX - XIII v.) Minsk.

Gurin, M. 2001a: Evolution of iron implements in Belarus. In: Archaeometallurgy


in The Central Europe III, 144 - 154.
BIBLIOGRAPHY 277
Gurin, M. 2001b: Metallographic examination of ancient iron half-finished products
in Belarus. In: Archaeology in The Central Europe III, 89 - 96.

Gurin, M. 2003: The welding is important element of blacksmith’s technology. In:


Archaeometallurgy in Europe, Milan, 229 - 238.

Gzelishvili, I. A. 1964: Zhelezoplavil’noye delo v drevney Gruzii. Tbilisi.

Hachulska-Ledwos, R. 1959/60: Wczesnośredniowieczny skarb żelazny z Mogily


pow. Kraków. Wiadomości Archeol. 26/3-4, 251 - 261.

Hadfield, R. 1912: Sinhalese iron and steel of ancient origin. JISI 85, 134 - 185.

Hagen, A. 1983: Norges Oldtid. Oslo.

Hallinder, P., Flyge, H., Randrup, J. 1986: The Iron Slag from Helgö. An Ar-
chaeological and Scientific Study. In: Excavations at Helgö X (A. Lundström, H. Carbe
eds). Stockholm, 131 - 151.

Hallinder, P., Haglund, K. 1978: Iron Currency Bars in Sweden. In: Excavations
at Helgö V:1. Stockholm, 30 - 58.

Hallinder, P., Tomtlund, J.-E. 1978: Rod-shaped blanks from Helgö. In: Excava-
tions at Helgö V:1. Stockholm, 59 - 80.

Hanemann, H. 1913: Metallographische Untersuchung einiger altkeltischer un an-


tiken Eisenfunde. Int. Zeitschrift f. Metallographie 4, 248 - 256.

Hanemann, H. 1921/1922: Metallographische Untersuchung einiger altkeltischen


Funde von der Steinsburg. PZ 13/14, 94 - 98.

Hanemann, H. 1930: Untersuchung eines eisernen Spitzbarrens aus der vorrömischen


Zeit. PZ 21 21, 271 - 274.

Hannack, G. 1930: Japanischer Damaststahl. Beiträge zur Geschichte der Technik


30, 87 - 90.

Hartwell, R. 1967: A Cycle of Economic Change in Imperial China: Coal and Iron
in Northeast China 75 – 1370. JESHO (Leiden), 10/1, 102 - 159.

Hauck, K. 1977: Wielands Hort. Die sozialgeschichtliche Stellung des Schmiedes in


frühen Bildprogrammen nach und vor der Religionswechsel. Antikvariskt arkiv vol. 64.
Stockholm.

Hauptmann. A., Mai, P. 1989: Chemische und mineralogische Untersuchungen


an Schlacken aus Colonia Ulpia Traiana. In: Spurenlese - Beiträge zur Geschichte des
Xantener Raumes (G. Precht, H. J. Schaller eds), 93 - 104.
278 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Hautmann, H., Morton, I. 1955: Metallographische Untersuchung eines von Hallstätter


Grabfeld stammenden Hufeisendolches. Jahrbuch d. Oberösterreichischen Musealvereines
100, 261 - 263

Hebenstreit, G. 2000: Historische Metallkunde des Eisens: Werkzeugaufbau und


Herstellungsmethode von mittelalterlichen Bergeisen und Bestimmung der Herkunft durch
dessen Schlackeneinschlüsse. Thesis. Montanuniversität Leoben.

Hegedüs, Z. 1960: Honfoglaláskori vastársyak és metallográfiai tanulaságai. Koházati


Lapok (Budapest) 1960/1.

Heimberg, U. 1997: Gesellschaft in Umbruch-Aspekten der Romanisierung - Gleich-


heitsnorme zwischen Rhein und Maas. Das Rheinische Landesmuseum Bonn - Berichte
aus der Arbeit des Museums 4/97, 79.

Heindel, I. 1993: Werkzeuge zur Metallbearbeitung des 7./8. bis 12./13. Jahrhundert
zwischen Elbe/Saale und Bug. Zeitschrift f. Archäologie 27, 337 - 379.

Hencken, H. 1968: Tarquinia, Villanovans and early Etruscans. Cambridge - Mas-


sachussets.

Henning, J. 1987: Eisenverarbeitungswerkstätten im unteren Donaugebiet zwischen


Spätantike und Frühmittelalter. Zeitschrift f., Archäologie (Berlin) 21, 59 - 73.

Henning, J. 1991: Schmiedegräber nördlich der Alpen - germanissches Handwerk


zwischen keltischen Tradition und römischem Einfluss. Saalburg Jahrbuch 46, 65 - 82.

Hermelin, E., Tholander, E., Blomgren, S. 1979: A Prehistoric Nickel-alloyed


Iron Axe, HM 13/2, 69 - 94.

Herrmann, F. R. 1969: Der Eisenhortfund aus dem Kastell Künzing. Saalburg


Jahrbuch 26, 129 - 141.

Hingley, R. 1991: Iron Age ‘Currency Bars’. The Archaeological and Sociological
Context. The Archaeological Journal 147 1990 (1991), 91 - 117.

Hingst, H. 1978: Vor- und frühgeschichtliche Eisenverhüttung in Schleswig-Holstein.


In: Eisen + Archäologie. Bochum, 67 - 71.

Hinton, D. A. 1993: A smith’s hoard from Tattershall Thorpe, Lincolnshire. Anglo-


Saxon England 22, 147 - 166.

Hjärthner-Holdar, E. 1993: Järnets och järnmetallurgins introduktion i Sverige.


Aun, vol. 16. Uppsala.

Hjärthner-Holdar, E., Kresten, P. Larsson, L. 1995: Currency bars and slags


BIBLIOGRAPHY 279
from the Trädsjördmästeren block - a reconnaissance, Uppland, town of Sigtuna. In: Ac-
tivity Report 1995, Geoarchaeological Laboratory, Uppsalla.

Hlubuček, K. 1953: Poslednı́ pilnı́káři a pilaři v Trhových Svinech. Doudlebský


archiv národopisný.

Holste, F. 1951: Hortfunde Südosteuropas. Marburg.

Holubowicz, W. 1954: Jak polscy kowale w XI w. imitowaly stal damasceńska̧.


Dawna Kultura (Wroclaw) 1954/3, 126 - 131.

Holubowicz, W. 1956: Opole w wiekach X - XII. Katowice.

Horedt, K. 1964: Die Verwendung des Eisens in Rumänien bis in das 6. Jh. v. u. Z.
Dacia N.S. VIII, 119 - 137.

Horstmann, D. 1989: Metallographische Untersuchung an Saxen mit gezahnter Zwis-


chenlage. Die Kunde N. F. 40, 209 - 211.

Horstmann, D. 1991: Metallographische Untersuchung an Saxen mit gezahnter Zwis-


chenlage. In: Westphal 1991, 364 - 365.

Horstmann, D. 1995a: Ferrum Noricum - Herstellung und Verbreitung. In: Gruben-


hunt & Ofensau. Vom Reichtum der Erde II. Klagenfurt, 277 - 282.

Horstmann, D. 1995b: Metallkundliche Unrersuchungen an Klingen von zwei römischen


Dolchen. Ausgrabungen und Funde in Westfalen-Lippe (München) 9/13, 111 - 135.

Hošek, J. 2001: Nickel-enriched bands in archaeological iron artefacts. In: Metallo-


graphy ‘01. Košice, 355 - 359.

Hošek, J. 2003a: Metallographic examination of the 10th to 12th centuries knives


fom the Přemyslid stronghold at Stará Boleslav (Bohemia). In: Archaeometallurgy in
Europe, 219 - 228.

Hošek, J. 2003b: Metalografie ve službách archeologie. Metallography in the service


of archaeology. Praha, Liberec.

Hošek, J. 2004: Blade construction at the 13th to 15th centuries Bohemian and Mora-
vian sickles - overview of the published research. In: Metallography 2004. Košice, 657 -
663.

Hošek, J., Mařı́k, J. 2004: Metallographic examination of the 10th century sword
from Kanı́n (Bohemia). In: Metallography 2004, Košice, 652 - 656.

Hošek, J. Prostřednı́k, Benešová, J. 1999: Kovářská dı́lna na hradě Trosky. The


smithy workshop in the castle Trosky. In: Z dějin hutnictvı́ 28, Rozpravy NTM 161.
280 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Praha, 25 - 35.

Høst-Madsen, L. Buchwald, V. F. 1999: The characterization and provenancing


of ore, slag and iron from the Iron Age settlement at Snorup. HM 33/2, 57 - 67.

Hübener, W. 1973: Die römischen Metallfunde von Augsburg-Oberhausen. Der Kat-


alog. Materialhefte zur Bayerischen Vorgeschichte (Kallmünz/Opf.) 28 1973.

Huggins, P. J., Huggins, R. M. 1973: Waltham Abbey: Excavation of Monastic


Forge and Saxo-Norman Enclosure 1972-73. Essex Archaeology and History 5, 127 - 184.

Hulme, E. W. 1933: Currency Bars and Water-Clocks. The Verdict of Archaeology


Reviewed. Antiquity 7, 61 - 72.

Huml, V., Pleiner, R. 1991: Die Schmiede im mittealalterlichen Prag. Die Eisen-
gegenstände aus der Notgrabung am Wenzelsplatz. Archaeologica Pragensia 11, 187 - 237.

Huml, V., Pleiner, R. 1993: Punch marks on high medieval iron artefacts. In:
Bloomery Ironmaking during 2000 Years III, 79 - 89.

Humert, J. 1991: Eine römische Strasse durch den südlichen Schwarzwald. Archäol.
Nachrichten aus Baden 45, 19 - 22.

Hvass, S. 1975: Das eisenzeitliche Dorf bei Hodde, Westjütland. Acta Archaeol.
(København) 46, 142 - 152.

Hvass, S. 1978: Die völkerwanderungszeitliche Siedlung Vorbasse, Mitteljütland. Acta


Archaeol. København 49, 61 - 111.

Iaroslavchi, E., Roşu, P. 1977: Cı̂tera noiaşezari dacica pe valea Cucuişului. Quelques
nouveax habitats daces dans la vallée de Cucuiş. Acta Musei Napocensis (Cluj) 14, 81 - 94.

IJzersterk: Experimenten ijzervinnen en smeden te Eindhoven 1993 (A. Boonstra ed.).


Eindhoven 1994.

Ilkjaer, J. 1976: Et bundt våben fra Vimose. Kuml 1975 (1976), 117 - 162.

Ilkjaer, J., Lønstrup, J. 1976: Nye udgravningen i Illerup. New excavations at


Illerup, Ådal. Kuml 1975 (1976), 94 - 115.

The Importance of Ironmaking. Technical Innovation and Change (G. Magnusson ed.).
Stockholm I 1995, II 1996.

Iron and Man in Prehistoric Sweden (H. Clarke ed.). Stockholm 1979.

The Introduction of Iron in Eurasia. Abstracts of papers of the Uppsala Conference 4


- 8 Oct. 2001 (S. Forenius ed.). Uppsala.
BIBLIOGRAPHY 281

Iron, Blacksmiths and Tools. Ancient European Crafts. Acta of the Instrumentum
Conference at Podsreda (Slovenia) in April 1996. Monographies instrumentum 12 (M.
Feugère, M. Guštin eds.). Montagnac 2000.

Jacobi, G. 1979: Drahtziehen der Latènezeit. Germania 57, 111 - 115.

Jančo, M. 2000: Germánska dielňa z Berouna. In: Sbornı́k Miroslavu Buchvaldkovi


(P. Čech, M. Dobeš eds.). Most, 107 - 110.

Jansová, L. 1992: Hrazany. Das keltische Oppidum in Böhmen III. Praha.

Jobey, G. 1968: A Radicarbon Date for the Palisaded Settlement at Huckhoe. Ar-
chaeologia Aeliana 46/4, 293 - 295.

Jockenhövel, A. 2001: Frühe Zangen. In: Festschrift für F.-R. Herrmann (S. Hansen,
V. Pingel eds.). Rahden/Westf., 91 - 102.

Johannsen, O. 1920: Chemische und metallographische Untersuchungen vorgeschicht-


licher Metallfunde. Stahl u. Eisen 40, 401 sq.

Jöns, H. 1997: Frühe Eisengewinnung in Joldelund, Kr. Nordfriesland. Ein Beitrag


zur Siedlungs- und Technikgeschichte Schleswig-Holsteins 1. Universitätsforschungen zur
prähistorischen Archäologie, vol. 40. Bonn.

Jöns, H., Wollschläger, B. 1999: Frühe Eisengewinnung in Südwestmecklenburg -


Ergebnis eines interdisziplinären Forschungsprojektes des Schwerpunktes ”Archäometallur-
gie”. Bodendenkmalpflege in Mecklenburg - Vorpommern, Jahrbuch 46. Lübestoff, 93 -
125.

Jouttijärvi, A., Lyngstrøm , H. 1990: Fire maend og deres jernknive. Vier Männer
und ihre Eisenmesser. Aarbøger for Nordisk Oldkyndighed og Historie (København))
1990, 59 - 67.

Kaczanowski, P. 1992: Importy broni rzymskiej na obszaru europeiskiego barbar-


icum. Imports of Roman armament on the territory of the European barbaricum. Kraków.

Kaenel, H. M. 1981: Ein Depotfund von 16 doppelpyramidförmigen Eisenbarren im


Schwadernamt. Arch. Suisse 4/1, 29 - 32.

Kamińska, J. 1968: Siedla̧tków, obronna siedziba rycerska z XIV wieku. Motte du


XIVe siècle de Siedla̧tków. Prace i Materialy Muzeum Archeol. i Etnograf. w Lodzi, ser.
arch. 15 - 88.

Kaszewska, E. 1981: Na bursztynowym szlaku. Z Otchlani Wieków 47/1-2, 26 - 29.

Keene, D. 1996: Metalworking in Medieval London: an Historical Survey. HM 30/2,


282 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
95 - 102.

Keesmann, I. 1985: Chemische und mineralogische Untersuchung von Eisenschlacken


aus der hallstattzeitlichen Siedlung von Niedererlbach. Archäeol. Korrespondenzblatt
15/3, 351 - 357.

Keesmann, I., Heege, A. 1990: Archäolmetallurgische Untersuchung an Mate-


rial der Ausgrabung 1984 am ”Steinbühl” bei Nörten-Hardenberg, Ldkr. Northeim.
Nachrichten aus Niedersachsens Urgesch. (Hildesheim) 59, 87 - 109.

Keesmann, I., Hellermann, B. 1989: Mineralogische und chemische Untersuchun-


gen an Schlacken von Morro de Mezquitilla. Madrider Mitteilungen (Mainz) 30, 92 - 117.

Keesmann, I., Hilgart, T. 1992: Chemische und mineralogische Untersuchung der


Schlacken von Manching. In: Ergebnisse der Ausgrabungen 1984 - 1987 in Manching (F.
Meier et al. eds.). Steinau, Stuttgart, 391 - 413.

Keesmann, I., Niemayer, H. G/, Golschani, F. 1983: Schlackenfunde von Toscanos.


Madrider Mitteilungem (Mainz) 24, 65 - 75.

Keesmann, I., Rieckhoff-Pauli, S. 1990: Eisenverarbeitung in der spätkeltischen


Siedlung Regensburg-Harting, Oberpfalz. Jahrbuch des Röm.-Germ. Zentralmuseums
(Mainz) 59, 553 - 556.

Keiling, H. 1972: Ein eiserner Amboss aus einem Grab der Jastorf-Kultur von Bod-
din, Kr. Hagenow. Ausgrabungen u. Funde 17/4, 176 - 179.

Keiling, H. 1987: Ein eiserner Hammer aus der frührömischen Kaiserzeit von Pleete,
Kr. Neubrandenburg. Ausgrabungen u. Funde 32/3, 136 - 139.

Khakhutaishvili, D. A. 1987: Proizvodstvo zheleza v drevney Kolchide. Tbilisi.

Kimmig, W. 1964: Seevölkerbewegung und Urnenfelderkultur. In: Studien aus Al-


teuropa I (R. von Uslar, K. Nau eds). Köln, Graz.

Kimmig, W., Gersbach, E. 1971: Die Grabungen auf der Heuneburg 1966 - 1969.
Germania 49, 29 - 91.

Kinder, J. 2003: Pattern-welded Viking-age sword blades - what can modern met-
allurgical investigation contribute to the interpretation of their forging technology? In:
Archaeometallurgy in Europe I, Milan, 239 - 246.

Kirpičnikov, A. N., Rjabinin, E. A., Petrenko, V. P. 1987: Das frühmittelalterliche


Ladoga im Lichte der neuesten Forschungen. Das Altertum 33/1, 54 - 61.

Kleemann, O. 1961: Stand der archäologischen Forschung über die eisernen Doppelpy-
ramiden(Spitz-)barren. Archiv f.d. Eisenhüttenwesen 32, 981 - 985.
BIBLIOGRAPHY 283

Kleemann, O. 1966: Der erste Fund vorgeschichtlicher Eisenbarren in Franken.


Mainfränkisches Jahrbuch f. Kunst u. Geschichte 18, 3 - 16.

Klein, J. 1972: A Greek Metalworking Quarter. Eight Century BC Excavations at


Ischia. Expedition (Phil.) 14/2, 34 - 39.

Klı́ma, B. jr 1985: Velkomoravská kovárna na podhradı́ v Mikulčicı́ch. Die grossmährische


Schmiede auf der Unterburg von Mikulčice. Památky archeol. (Praha) 76, 428 - 455.

Klingenbiel, H., Hundeshagen, H. 1957. Der Schmied am Amboss. Berlin.

Koch, U. 1984: Der Runde Berg bei Urach V: Die Metallfunde der frühgeschichtlichen
Perioden aus den Plangrabungen 1967 - 1981. Heidelberg.

Koch, U. 1988: Ein Depotfund vom Runden Berg: Gerätschaften eines alamanischen
Wirtschaftbetriebes der Terassensiedlung. Archäeol. Korrespondezblatt 18/2, 205 - 208.

Koch, U. 1997: Damaszieren von Waffen. In: Europäische Technik im Mittelalter 800
bis 1400 - Tradition und Innovation. (U. Lindgren ed.). Berlin, 217 - 220.

Kokowski, A. 1981: Pochówki kowali w Europie w IV w. p.n. e. do VI w. n. e.


Schmiedegräber in Europa vom IV. Jh. v.d.Z. bis zum VI. Jh. n.d.Z. Archeologia Polski
26/1, 191 - 218.

Kola, A., Wilke, G. 1975: Produkcja grotów beltów do kuszy w średniowieczu w


świetle wspólczesnych prób experimentalnych. Uwagi o odkryciach na grodzisku póżno
śreniowiecnym w Sloszewach, pow. Brodno. Die Produktion der Armbrustpfeilspitzen
im Mittelalater im Lichte der heutigen Experimentalversuche. Acta Univ. N. Copernici
(Toruń), Arch. V, 162 - 181.

Kolchin, B. A. 1949: Opyt metallograficheskogo issledovaniya drevnerusskikh vecht-


chey. KS 1949/30, 42 - 45.

Kolchin, B. A. 1953: Tchornaya metallurgiya i metalloobrabotka v drevney Rusi.


MIA 32, Moskva.

Kolchin, B. A. 1959: Zhelezoobrabatyvayuchtcheye remeslo Novogoroda Velikogo.


In: Trudy novgorodskoy ekspedicii II. MIA 65, Moskva, 7 - 130.

König, P. 1986: Schmiedehandwerk in Haithabu. Zur Aussage der Eisenschlacken-


funden. Thesis, Kiel.

Korolev, A. B., Khlebnikova, T. A. 1960: K voprosu o chernoy metallurgii u


volzhskikh Bolgar. MIA 80, Moskva, 159 - 168.

Koryakova, L., Beltikova, G., Kuzminich, S. V. 2001: The Introduction of Iron


284 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Technology in Central-Northern Eurasia (Eastern Europe, Ural and Western Siberia). In:
The Introduction of Iron in Eurasia, 20 - 21.

Kostrzewski, J. 1949: Pradzieje Polski. Poznań.

Kostrzewski, J. 1953: Wytwórczosć metalurgiczna w Polsce od neolitu do wczesnego


okresu żelaznego. La production métallurgique en Pologne depuis la néolithique jusq’au
premier Âge du fer. Przegl. Archeol. 9 1950/1953, 177 - 213.

Kotarba, J., Leblanc, J.-C. 1998: L’habitat et l’atelier métallurgique du site des
Coudoumines III à Caraman (Pyrénées Orientales). In: Recherches sur l’économie du fer,
182 - 209.

Kozák, K. 1981: Az Égri és Pécsváradi vár (kolostor) XV - XVI századi kemencék.
15th and 16th century furnaces in the Eger Castle and Pécsvárad monastery. In: Iparégeszet
I 1980 (1981), 133 - 134.

Krämer, W. 1989: Das eiserne Ross von Manching. Fragmente einer Mittellatène
Pferdeplastik. Germania 67/2, 519 - 239.

Krajı́c, R. 1993: Středoěká kovárna v Sezimově Ústı́ - Novém Městě. Eine Schmiedew-
erkstatt in Sezimovo Ústi - Neustadt. Archaelogica Historica 18/93, 391 - 417.

Krajı́c, R.: 2003: Sezimovo Ústı́ - Archeologie středověkého poddanského města 3.


Kovárna v Sezimově Ústı́ a analýzy výrobků ze železa. Sezimovo Ústı́ - Archäolgie der
mittelalterlichen Untertanenstadt. Die Schmiede in Sezimovo Ústı́ und Analyse der Pro-
dukte aus Eisen I-II. Praha, Sezimovo Ústı́, Tábor.

Krajı́c, R., Kukla,, Z., Nekuda, P. 1997: Středověký meč z Mstěnic. Das mittelal-
terliche Schwert von der Wüstung Mstěnice. In: Z pravěku do středověku. Brno, 250 - 258.

Kromer, K. 1959; Das Gräberfeld von Hallstatt. Florenz.

Kromer, K. 1964: Vom frühen Eisen und reichen Salzherren. Die Hallstattkultur in
Österreich. Wien.

Kronz, A. 1997: Phasenbeziehungen und Kristallisationsmechanismen in fayalitischen


Schmelzsystemen - Untersuchungen an Eisen- und Buntmetallschlacken. Dissertation,
Mainz.

Krupkowski, A., Reyman, T. 1953: Badania metaloznawcze nad przekutym pól-


fabrykatem żelaza z Witówa, pow. Pińczów, i żużlem dymarkowym z Igolomi, pow.
Miechów. Metallurgical Investigation of Iron Bars Found at Witów and slag found at
Igolomia. Sprawozdania Państw. Muz. Archeol. 5/1-2, 48 - 65.

Kudrnáč, J. 1984: A high medieval smithy at a gold mine of Čelina near Přı́bram,
South-Central Bohemia. Archeol rozhledy (Praha) 36, 683.
BIBLIOGRAPHY 285

Kudrnáč, J. 1992: Recherches archéologiques sur la production médievale d’or en


Bôheme. In: Les techniques minières de l’antiquité au XVIIIe siècle. Paris, 299 - 313.

Kutchera, S. 1977: Kitayskaya arkheologia 1965 - 1979 gg. Moskva.

Kuzminykh, S. V. 2001: Iron in the Bronze Age Cultures of North Eurasia. The
Introduction of Iron in Eurasia. Conference abstracts. Uppsala, 21 - 22.

La Salvia, V., Mihok, L’., Pribulpvá, A. 2001: Medieval iron metallurgy in town
Cencelle, Italy. In: Archaeometallurgy in The Central Europe III, 155 - 174.

La sidérurgie ancienne de l’Est de la France dans son contexte européen - Archéologie


et archéometrie (M. Mangin ed.). Actes du colloque de Besançon. Paris 1994.

Lang, J. 1988: Study of the metallurgy of some Roman swords. Britannia 19, 199 -
216.

Lang. J. 1991: A Dirk from Cyprus. In: From Bloom to Knife, 93 - 96.

Larsson, L., Hjärthner-Holdar, E. 2000. Pornullbacken. In: Activity Report 1998


(Uppsala 2000), 13 - 15.

Laumann, H. 1985: Ein spätlatènezeitlicher Schmiedeplatz von Neunkirchen-Zeppen-


feld, Kr. Siegen-Wittgenstein. Ausgrabungen u. Funde in Westfalen-Lippe 3, 49 - 57.

Le fer (M. Mangin, ed.), with contributions by F. Dabosi, C. Domergue, P. Fluzin, M.


Leroy, M. Mangin, P. Merluzzo, A. Ploquin, V. Serneels. Editions Érrance, Paris 2004.

Lebeaupin, D. 1998: Atelier des forgerons et témoins dispersés du travail du fer à


Lattes (Hérault) (IVe s. av. - Ier s. ap. J.-C.). In: Recherches sur l’économie du fer, 80
- 95.

Leblanc, J.-C. 1997: Caractérisation d’une activité spécialisée: des forgerons-charrons


au premier siècle après J. -C. (Cité Judiciaire, Bordeaux). In: Mélanges Claude Domer-
gue, Pallas 46 (Presses Universitaires du Mirail). Aix-en-Provence, Montpellier, Perpi-
gnan, 251 - 263.

Leciejewicz. L., Losiński, W., Tabaczyńska, E. 1961: Kolobrzeg we wczesnym


średniowieczu. Wroclaw.

Lenz, E. von 1908: Über Damast. Zeitschrift f. hist. Waffenkunde 4 1906/1908, 132
- 136.

Leoni, M. 1955: Archeologia e metallografia. Sibrium 2, 25 - 42.

Leoni, M. 1975: Tradizione e realtà delle spade Galliche. Sibrium 22 (1973/1975),


286 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
105 - 125.

Leoni, M., Panseri, C. 1961: La tecnologia del ferro presse gli Etruschi. Studi Etr-
uschi X 29. Firenze, 235 - 234.

Lepage, L. 1966: Étude d’un fer de javelot hallstattien de Vest-La-Gravelle, Marne.


Mémoires de la Société d’agriculture, commerce, science et arts du département de la
Marne LXXXI (Châlons sur Marne), 7 - 11.

Len’kov, V. D. 1974: Metallurgiya i metalloobrabotka u Churchzheney v 12 veke (po


matarialam Shayginskogo Gorodichtcha). Novosibirsk.

Les métaux antiques: travail et restauration. Actes du colloque de Poitiers 28-30 sept.
1995. (G. Nicolini, N. Dieudonnée-Glad eds). Monographies instrumentum 6, Montagnac
1998.

Levinsen, K. 1983: Jernets introduktion i Danmark. The Introduction of Iron to


Denmark. Kuml (Københvn) 1982/83, 153 - 161.

Leube, A. 1996: Ein frühgeschichtliches Schmiedegerätdepot von Jütchendorf bei


Zossen im Teltow. Acta Praeh. et Archaol. 28, 59 - 93.

Li Chung 1979: Studies in the iron blade of a Shang dynasty at Kuo-Ch’eng, Hopei,
China. Ars Orientalis 11, 259 - 289.

Liestøl, A. 1961: Blodrefill og Mål. Viking 1961, 71 - 98.

Lietzmann, K.-D., Schlegel, J., Hensel, A. 1984: Metallformung. Geschichte,


Kunst, Technik. Leipzig.

Limet, H. 1976: Les origines du travail du fer à Mésopotamie ancienne. Présence 92,
3 - 16.

Linderoth, B. 1991: L’Anse aux Meadows - Gateway to Vinland. In: The Norse of
the North Atlantic. Acta Archaeol. København 61, 1990, 166 - 197.

Livadefs, C. J. 1956: The Structural Iron of the Parthenon. JISI 182, 49 - 66.

Loehberg, K. 1969: Untersuchungen an Eisenfunden aus Limeskastellen. Saalburg


Jahrbuch 26, 142 - 146.

Losiński, W. 1959: Kowalstwo we wczesnośredniowiecznym Kolobrzegu. Prace Ko-


misji Archeol. PTPPN 4/1-1, 9 - 56.

Lu Da 1965: Die uralte Technik der Eisenherstellung in China. In: Vita pro Ferro.
Festschrift Durrer. Schaffhausen, 63 - 70.
BIBLIOGRAPHY 287
Lüdemann, K.-F., Ebert, R., Schirmer, W. 1962: Ergebnisse der Untersuchung
einiger vor- und frühgeschichtlichen Eisen- und Schlackenfunde. Ausgrabungen u. Funde
7/1, 12 - 18.

Lunegov, J. A. 1956: Redikarskiy mogil’nik. KS 57, 115 - 123.

Lyngström, H. 1997: In the borderland of archaeology - experimental forging. In:


Early Iron Production - Archaeology, Technology and Experiments, 27 - 35.

MacCormick, A. 1996: Metalworking in medieval Nottingham, HM 32/2, 103 - 110.

Mack, J., McDonnell, G., Murphy, S., Andrews, P. Wardley, K. 2000: Liquid
steel in Anglosaxon England. HM 34/2, 87 - 96.

Macrea, M., Rusu, M. 1960: Der dakische Friedhof von Porolissum und das Prob-
lem der dakischen Bestattungsbräuche in der Spätlatènezeit. Dacia 4, 201 - 229.

Maddin, R. 1985: The technology of ironmaking. In: Medieval Iron in Society I, 127
- 157.

Maddin, R., Hauptmann, A., Baatz, D. 1991: A Metallographic Examination of


Some Iron Tools from the Saalburgmuseum. Saalburg Jahrbuch 66, 5 - 23.

Maddin, R., Hauptmann, A., Weisgerber, G. 1995: The Metallurgy of Roman


Iron Tools. In: Paléometallurgie du fer & Cultures, 525 - 535.

Madsen, H. G. 1981: Tuyeres. In: Excavations at Helgö VII. Stockholm, 95 - 105.

Maes-Diop, L.-M. 2002: Bilan des datations des vestiges anciens de la sidérurgie en
Afrique. L’enseignement qu s’en dégage. In: Aux origines de la métallurgie du fer en
Afrique, 189 - 193.

Magyar, K. 1999: The tools of the 17th century forge in oppidum Segest. In: Tradi-
tions and Innovations, 230 - 240.

Maier, R. A. 1986: Eisenäxte von altslawisch-grossmährischen Typ aus dem Inn bei
Töging im Museum Altötting (Oberbayern). Germania 64/1, 180 - 183.

Malinowski, T. 1953: Narzȩdzia kowalskie okresu póżnolateńskiego i rzymskiego w


Polsce. Przegl. Archeol. IX/2-3, 258 - 280.

Mangin, M., Fluzin, P. 2002: La petite sidérurgie en contexte rural: Les forges du
Haut-Auxois et de ses margins Lingonnes (Côte d’Or). In: Cultivateurs, éleveurs et arti-
sans dans les campagnes de Gaule romaine. Matières premières et produits transformés.
VIe Colloque AGER. Amiens, 141 - 156.

Mangin, M., Fluzin, P., Courtadon, J.-L., Fontaine, M.-J. 2000: Forgerons et
288 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
paysans des campagnes d’Alésia (Haut-Auxois, Côte d’Or), 1er siècle avant - VIIIe siècle
aprs̀ J. C. Monographies CRA 22, Paris.

Mangin, M., Fluzin, P. Ratel, D. and R. 2002: La forge du sanctuaire rural


gallo-romain d’Ancernant (Côte d’Or). Revue archéologique d’Est 51 2001/2002, 251 -
298.

Manning, W. H. 1964: A Roman Hoard of Ironwork from Sandy, Bedfordshire. Bed-


fordshire Archaeol. Journal 2, 50 - 57.

Manning, W. H. 1965: A Relief of Two Greek Freedmen. The British Museum


Quarterly XXIX/1-2, 25 - 28.

Manning, W. H. 1966: A Hoard of Romano-British Ironwork from Brampton, Cum-


berland. Transactions of the Cumberland and Westmoreland Antiquarian and Archaeo-
logical Society 66 N.S/, 1 - 35.

Manning, W. H. 1972: Ironwork hoards in Iron Age and Roman Britain. Britannia
III, 224 - 250.

Manning, W. H. 1977: Blacksmith’s Tools from Waltham Abbey, Essex. In: Aspects
of Early Netallurgy (W. A. Oddy ed.). Symp. British Museum, 22 - 23 April 1977, 87 - 96.

Marghitan, L., Andrit.oiu, I. 1971: Muzeul arheologic Deva (Cluj) 35, 15.

Marić, Z. 1979: Depo pronaden u ilirskomu gradu Daors (2. st. pr. n. e.). Glasnik
Zemaljskog muzeja Sarajevo, Archaologija N.S. 33 1978 (Sarajevo 1979) 23-8, 115.

Marichal, R. 2000: Outillage antique de Ruscino (Château Rousillon, Pyrénées-


Orientales, F). In: Iron, Blacksmiths and Tools, 139 - 168.

Martens, I. 1979: Jerndepotene - moen aktuelle problemstillingar. Eisendepots -


einige aktuelle Fragenstellungen. In: Jern og Jernvinna som middelalder i Norge. Sta-
vanger, 59 - 69, 110.

Martens, I. 2002: Smeden og hans produkter i norsk vikingtid. The blacksmith and
his products. In: UKM - En manyfoldig forskningsinstitusjon (E. Høigård ed.). Oslo, 173
- 185.

Martin, Th., Ruffat, H. 1998: Un dépôt de lingots de fer du début de La Tène III
à Montans (Tarn). In: In: Recherches sur l’économie du fer, 110 - 115.

Maryon, H. 1960: Pattern-welding and Damascening of Sword-blades, Parts 1 - 2.


Studies in Conservations 5/1 - 2, 25 - 77.

Mathieu, S., Mignot, P., Plumier, J. 1994: Structures associées à des scories de
forge sur quelques sites romains en Belgique. In: La sidérurgie ancienne de l’Est de la
BIBLIOGRAPHY 289
France, 141 - 142.

Maufras, O., Fabre, L. 1998: Une forge tardive (fin IVe - Ve s.) sur le site de la
Ramière (Roquemaure, Gard). In: Recherches sur l’économie du fer, 210 - 221.

Maurer, H.-P. 1993: Archäometallurgische Untersuchungen and Schlacken und Eisen-


funden de latènezeitlichen Schmiedewerkstatt Kundl-Lus und aus dem Oppidum von
Manching. Archäol. Korrespondenzblatt 23/3, 313 - 325.

Mazur, A., Nosek, E. 1972: Wczesnośredniowieczne noże dziwerowane z Wroclawia.


Early medieval ornamented knives from Wroclaw. Kwartalnik Hist. Nauki i Techniki
17/2, 291 - 304.

Mazur, A., Nosek, E. 1989: Investigation of the hoard of currency bars from Kraków.
In: Archaeometallurgy of Iron 1967 - 1987, 429 - 435.

McDonnell, G. 1991: A Model for the Formation of Smithing Slags. In: From Bloom
to Knife, 23 - 26.

McDonnell, J. G. 1992: Ironworking. In: Anglo-Scandinavian Ironwork from 16-22


Coppergate. The Archaeology of York - The Small Finds 17/6. London, 471 - 736.

McLees, C. 1996: Itinerant craftsmen, permanent smithies and the archbishop’s mint:
the character and context of metalworking in medieval Trondheim. HM 30/2, 121 - 135.

Medieval Iron in Society I - II, Symposium at Norberg. Stockholm 1985.

Meier, D., Reichstein, J. 1984: Eine wikingerzeitliche Siedlung westlich von Kosel,
Kreis Rendsburg-Eckenförde (La 117). Offa 41, 113 - 168.

Merkel, J., Barrett, K. 2000: The adventitious production of iron in the smelting
of copper’ revisited: metallographic evidence against a tempting model. HM 34/2, 59 - 66.

Mesterházy, K. 1990: Münzdatierte spätkaiserzeitliche Gerätfunde aus Tadaj. Alba


Regia 24, 53 - 66.

Metallgewinnung und -verarbeitung in der Antike (Schwerpunkt Eisen). (H. Friesinger,


K. Pieta, J. Rajtár eds). Nitra.

Metallography 2004. Acta Metallurgica Slovaca 10/1. Special issue, 12th International
Symposium on Metallography. Košice.

Michálek, J. 1999: Keltský poklad z Bezdědovic na Blatensku. Eisengerätfund der


Latènezeit aus Bezdědovice bei Blatná, Südböhmen. Metallkundliche Analysen by R.
Pleiner, B. Novotná, 9 - 16. Blatná. Strakonice.

Mikheyev, V. K. 1966: Do pytannya po remisniche vyrobnictvo saltovs’koy kultury.


290 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Visnik Kharkovskogo universiteta 17, 91 - 95.

Miglbauer, R. 1988: Ein römischer Verwahrfund aus Wels, OÖ. Bayerische Vorge-
schitsblätter 53, 287 - 292.

Mihok, L’., Pribulová, A., Unger, J. 1997: Kováčská dielňa na stredovekom hrade
Lelekovice. Z dějin hutnictvı́ (Nat. Tech. Mus. Praha) 26, 31 - 57.

Mihok, L’., Pribulová, A., Unger, J. 1999 (et. al. 1999c): Smithy production
in Medieval Castle Lelekovice, south Moravia. In: Archaeometallurgy in The Central
Europe, 115 - 131.

Mihok, L’., Pribulová, A., Fröhlichová, M., Kotigorochko, V. M. 1999 (et


al. 199a): Research of Latène and Roman slags from smelting and working of iron from
upper Tisza Region. In: Archaeometallurgy of The Central Europe, 16 - 26.

Mihok, L’., Senn, M. 1995: Structure of smithy slags from Roman period in Neften-
bach and Dietikon, Switzerland. In: Metallography ’95, Košice, 418 - 421.

Mihok, L’., Pribulová, A., La Salvia, V. 1999 (et al. 1999b): Smithy production in
medieval town Leopoli-Cencelle. In: Archaeometallurgy in The Central Europe, 192 - 212.

Mihok, L’., Pribulová, A. 2003: Metallography of Avar Blacksmiths products: Ar-


chaeometallurgy in Europe 1, Milan, 181 - 188.

Mihovilić, K. 1999: Nalezi prahistorijskich ostava na području Istre. Vorgeschichtli-


che Hortfunde in Istria. Archeol. vesnik 42, 207 - 217.

Minto, A. 1943. Populonia, Firenze.

Miroššayová, E. 1980: Depot železných predmetov z Nižnej Myšle. Depot von Eisen-
gegnständen aus Nižná Myšla. Slov. archeol. 28/2, 383 - 394.

Modin, S., Pleiner, R. 1978: The metallographic examinations of lock, keys and
tools. In: Excavations at Helgö V: 1. Stockholm, 81 - 109.

Modin, S., Thålin, L. 1969: Metallographic aspect of prehistoric spear forging.


Jernkontorets Ann. 153/5, 231 - 239.

Mohen, J.-P. 1978: L’outilage des bronziers de l’âge du Bronze en France. In:
Journées de Paléomeallurgie. Compiègne, 107 - 124.

Mohen, J.-P. 1980: L’âge du fer en Aquitanie du VIIIe au IIIe siècle avant Jésus
Christ. Mém. de la Société Préhistorique Française vol. 14, Paris.

Mohen, J. P. 1990: Métallurgie préhistorique. Introduction à la paléometallurgie.


Paris, Milan, Barcelone, Mexico.
BIBLIOGRAPHY 291

Money, J. H. 1974: Iron Age and Romano-British iron-working site in Minepit Wood,
Rotherfield, Sussex. HM 8/1, 1 - 20.

Montanarchäologie in Europa. Berichte zu einem Kolloquium (H. Steuer, U. Zimmer-


mann eds). Sigmaringen 1993.

Moosleitner, F., Urbanek, E. 1991: Das Werkzeugdepot einess keltischen Grob-


schmiedes von Nikolausburg bei Golling, Land Salzburg. Germania 69, 63 - 78.

Morton, T. 1954: Analysen von Eisenschlacken und Eisenwerkzeugen aus der römischen
Niederlassung in der Lahn und vom Grabfeld in Hallstatt. Jahrbuch des Oberösterr.
Musealverines (Linz) 99, 177 - 180.

Morton, F., Hautmann, H. 1957: Chemische Analysen und metallographische Un-


tersuchungen von Eisenerzen und Eisengegnständen von der Dammwiese und der römischen
Niederlassung in der Lahn. Jahrbuch des Oberösterr. Musealvereines (Linz) 102, 133 -
135.

Moschkau, R. 1962: Eisernes Gebrauchsgerät der Spät-Latènezeit als Grabfund von


Leipzig-Thekla. Ausgrabungen u. Funde 7/2, 83 - 88.

Możdzioch, S. 1990: Organizacja gospodarstwa państwa wczesno-piastowskiego na


Śla̧sku - studia archeologiczne. Wirtschaftliche Organisation der frühen Piastenstaates
(10. - 12. Jh.) im Lichte der archäologischen Quellen. Wroclaw, Warszawa, Kraków.

Müller, D. W. 1984: Brückenschlag über Jahrhunderte. Wegweiser durch die Schau-


sammlung des Landesmuseums f. Vorgeschichte Halle (Saale). Berlin.

Müller, F. 1990: Der Massenfund von Tiefenau bei Bern. Antiqua 20. Basel.

Müller-Karpe, A and M. 1977: Neue latènezeitliche Funde aus dem Heidetränk-


Oppidum in Taunus. Germania 55, 33 - 63.

Müller-Wille, M. 1977: Der frühmittelalterliche Schmied im Spiegel skandinavischer


Grabfunde. Frühmittelalterliche Studien (Münster) 11, 127 - 211.

Müller-Wille, M. 1983: Der Schmied im Spiegel archäologischer Quellen. Zur Aus-


sage von Schmiedegräbern der Wikingerzeit. In: Das Handwerk in vor u. frühgeschichtlicher
Zeit II. Göttingen, 216 - 260.

Munksgaard, E. 1984: A Viking Age Smith, his Tools and his Stock-in-Trade. Offa
41, 85 - 89.

Musgrave, M. P. 1993: Examination of Roman iron artefacts from Caerwent Roman


city. Bull. of the Board of Celtic Studies 40, 212 - 239.
292 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Mutz, A. 1960: Römische Fenstergitter. Jahrbuch der Schweiz. Gesellschaft f.
Urgeschichte 48, 107 - 112.

Mutz, A. 1968: Römische Eisenwerkzeuge aus Augst. Technologische Untersuchun-


gen und Vergleiche mit modernen Parallelen. In: Festschrift Laur-Belart (Basel), 155sq.

Mutz, A. 1976: Römisches Schmiedehandwerk, Augster Museumshefte 1.

Nägler, T. 1979: Ein mittelalterlicher Werkzeugfund aus Pretai. Forschungen zur


Volks- u. Landeskunde 22/2, 24 - 29.

Narysy starodavnoy istoriy Ukrayinskoy RSR. Kiev 1957.

Naumann, F. K. 1957: Untersuchung eines eisernen luristanischen Kurzschwertes.


Archiv f. d. Eisenhüttenwesen 28, 575 - 581.

Naumann, F. K. 1964: Untersuchungen alter Fundstücke von den Ausgrabungen am


Magdalensberg in Kärnten. Archiv f. d. Eisenhütenwesen 35/6, 405 - 502.

Naumann, F. K. 1965: Die Untersuchung mittelalterlichen Eisenfunde aus Bargen


im Kanton Schaffhausen. In: Vita pro Ferro, Festschrift Durrer. Schaffhausen, 195 - 210.

Naumann, F. K. 1971: Metallkundliche Untersuchungen an drei wikingerzeitlichen


Zieheisen aus Haithabu. In: Berichte über die Ausgrabungen in Haithabu (Neumünster)
vol. 5, 84 - 99.

Navasaitis, J. 2003: Lietuviška geležis. Kaunas.

Needham, J. 1964: The Development of Iron and Steel Technology in China. Cam-
bridge.

Neubeck, A., Schaaber, O. 1965: Metallographic studies of Roman-Noric iron ex-


cavations. Freiburger Forschungshefte B 111, 7 - 26.

Neuburger, A. 1919: Die Technik des Altertums. Leipzig.

Neumann, B. 1928: Römischer Damaststahl. Archiv f. d. Eisenhüttenwesen 1


2927/28, 241 - 244, 286.

Neumann, B., Klemm, H. 1949: Metallographische Untersuchung von eisernen


Dübeln und Klammern aus dem 2200 Jahren alten Artemis-Tempel von Magnesia am
Mäander. Archiv f. Metallkunde 3, 333 - 335.

Nihlén, J. 1939: Äldre Järntillverkning i Sydsverige. Uppsala.

Nikolov, B. 1970: Kolektivna nakhodka ot zhelezni predmeti ot khalschatskata epokha


do gr. Krividil, Vrachanski okr’g. Arkheologiya (Sofia) 1971/1, 51 - 57.
BIBLIOGRAPHY 293

Nikolskaya, T.N. 1957: Drevnerusskoye selichtche Lebedka. L’ancien village russe


Lebedka. Sov. arkh. 4/1957, 176 - 197.

Nikulka, F. 2000: Zur Genese der Eisenmetallurgie in Nordwestdeutschland. Archäol.


Mitteilungen aus Nordwestdeutschland 23, 59 - 105.

Nørbach, L. Chr. 1998: Ironworking i Denmark. From the Late Bronze Age to the
Early Roman Iron A. Acta Archaeol. (Copenhagen) 69, 53 - 75.

Nordahl, E. 1988: Reykjavik from the Archaeological Point of View. Uppsala.

Nosek, E. 1966a: Badania metalograficzne wȩdzidla z Chelmca, pow. Nowy Sa̧cz.


Wiad. Archeol. 32/1-2, 12 - 15.

Nosek, E. 1966b: Niektore zabytki żelazne z terenu Gór Świȩtokrzyskich w świetle


badań metaloznawczych. Materialy Archeol. (Kraków) 7, 179 - 184.

Nosek, E. 1968. Czternastowieczna kużnia w Siedla̧tkowie. A Fourteenth Century


Forge at Siȩdlatków. Prace i Mat. Muzeum Archeol. i Etnograf. w Lodzi, ser. arch. 15,
95 - 132.

Nosek, E. 1991: Forging of high phosphorus iron. In: From Bloom to Knife, 67 - 70.

Nothdurfter, H. 1979: Die Eisenfunde von Sanzeno im Nonsberg. Mainz am Rhein.

Novotný, B. 1963: Výzkum velkomoravského hradiště ”Pohansko”: u Nejdku na Led-


nickém ostrově. Die Erforschung des grossmährischen Burgwalles ”Pohansko” bei Nejdek
auf der Lednice Insel. Pam. archeol. (Praha), 54, 3 - 40.

Novotný, B. 1969: Depots von Opfersymbolen als Reflex eines Agrarkultes im Grossmähren
und in wikingerzeitlicher Skandinavien. Pam. archeol. (Praha) 60, 197 - 227.

Nylén, E. 1981: Von Bronze zu Eisen in Schweden, eine kulturgeographische Analyse.


In: Frühes Eisen in Europa, 89 - 100.

Ocherki po istorii drevney zhelezoobrabotki v Vostochnoy Evrope. Essays on History


of Ancient Ironworking in Eastern Europe. Moskva 1997.

Ohlhaver, H. 1939: Der germanische Schmied und sein Werkzeug. Leipzig.

Olive, Chr., Ugolini, D. 1998: Le travail du fer à Béziers (Hérault). In: Recherches
sur l’économie du fer, 76 - 79.

Oppenheim, M. von 1950: Tell Halaf II. Berlin.


294 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Orengo, L., Bonnon, J.-M., 2000: L’emploi des blocs-tuyères dans les forges an-
tiques du centre de la Gaule (Auvergne, Lyonnais et Forez au Deuxième âge du Fer et à
l’époque romaine). Découvertes archéologiques et expérimentation. In: Iron, Blacksmiths
and Tools, 121 - 136.

Ottaway, P. 1992: Anglo-Scandinavian Ironwork from 16-22 Coppergate. The Ar-


chaeology of York - The Small Finds 17/6. London, 453 - 736.

Overbeck, J. 1856: Pompeji in seinen Gebäuden, Alterthümern und Kunstwerken.


Leipzig.

Pachkova, S. P., Gopak, V. D. 1981: Zalizoobrobne remeslo na horodichtchi


Hrynchuk. Arkheolohiya (Kyiv) 36, 54 - 66.

Paléometallurgie du fer & Cultures. Symp. Internat. du CPSA (P. Benoit, P. Fluzin,
eds.). Paris 1995.

Panseri, C. 1957: La tecnica di fabricazione delle lame di acciaio presso gli antichi.
Associazione Italiana di Metallurgia (Milano), 19 - 33.

Panseri, C., Leoni, M. 1966: The manufacturing technique of Etruscan weapons.


Examination of an iron spear-head (IV century B. C.) from Montefiascone. Metallurgia
Italiana 58, 381 - 389.

Paret, (O.) 1934: Arbeitsgebiet des Landesamtes für Denkmalpflege Stuttgart. Ger-
mania 18/1, 62.

Pašalić, E. 1954: O antičkom rudarstvu u Bosni i Hercegovini. L’exploitation des


mines dans l’antiquité en Bosnia-Herzégovina. Glasnik Zemaljskog muzeja u Sarajevo -
arch. NS 9, 47 - 72.

Passelac, M. 1998: Installations pour le travail du fer dans le Vicus Eburomagus


(Bram, Aude). With contributions by L. Chabal, J.-C. Leblanc. In: Recherches sur
léconomie du fer, 129 - 141.

Paulı́k, J. 1970: Najstaršie hromadné nálezy železných predmetov na Slovensku


(Keltské depoty železiarských výrobkov v Plaveckom Podhradı́). Die ältesten Hort-
funde von Eisengegenständen in der Slowakei (Keltische Depote von Eisenerzeugnissen
in Plavecké Podhradie). Zbornı́k Slov. nár. muzea (Bratislava) 64, História 10, 25 - 83.

Pellecuer, Chr. 1998: Le travail du fer dans la villa des Prés Bas à Loupian (Hérault).
In: Recherches sur l’économie du fer, 166 - 174.

Perdrizet, P. 1908: Fouilles de Delphes V. École Français d’Athènes. Paris.

Pérez-Suñé, J. Ma., Revilla Calvo, V., Gómez Sánches, J., Simóm Arias,
Marsal Astort, M., Plana Llevat, F. 1998: Functión de la siderurgia en la Cataluña
BIBLIOGRAPHY 295
romana: In: Recherches sur l’économie du fer, 222 - 250.

Pertlwieser, M. 1970 (1972): Die hallstattzeitliche Höhensiedlung am Waschenberg


bei Bad Wimsbach/Neydharting, Politischer Bezirk Wels, Oberösterreich. Jahrbuch d.
Oberösterr. Musealvereines (Linz) 115/1 1970 (1972), 37 - 70.

Peschel, K. 1980: Der Hortfund von Leipzig-Wahren. Arbeits- u. Forschungsberichte


z. Sächs. Bodendenkmalpflege 23, 35 - 52.

Petrescu-Dimboviţa, M. 1978: Scurtă istoria a Daciei preromane. Iaşi.

Petrie, W. M. F. see Flinders Petrie.

Photos, E. 1989: The question of meteoric iron versus smelted nickel-rich iron: ar-
chaeological evidence and experimental results. World Archaeology 20/3, 403 - 421.

Piaskowski, J. 1954 (1956): Badania przedmiotów metalowych z grodziska w Za-


wadzie Lanckorońskiej (pow. Brzesko). Kwartalnik Hist. Nauki i Techniki 2/1954 (1956),
375 - 387.

Piaskowski, J. 1958a: Badania metaloznawcze przedmiotów żelaznych z kurhanów


okresu rzymskiego we wsi Szwajcaria, pow. Suwalki. Wiadomości Archeol. 25/1-2, 53 -
71.

Piaskowski, J. 1958b: Metaloznawcze badania zabytków archeologicznych z Wycia̧ż,


Igolomi, Jadownik Mokrych i Piekar. Studia z Dziejów Gornictwa i Hutnictwa II, 7 - 98.

Piaskowski, J. 1959: Cmentarzysko z XI wieku w Lutomiersku pod Lodzia̧. In: A.


Nadolski, A. Abramowicz, T. Poklewski. Cmentarzysko z XI wieku w Lutomiersku pod
Lodzia̧, Lódż, 110 - 31.

Piaskowski, J. 1959b: Metaloznawcze badania wyrobów żelaznych z cmentarzysk


cialopalnych Wielkopolski z okresu halsztackiego. Fontes Archeol. Posnanienses 10, 202
- 228.

Piaskowski, J. 1959c: Technika wczesnośredniowiecznych wyrobów dziwerowanych w


świetle novych badań. Przegla̧d Mechaniczny 18/15, 495 - 499.

Piaskowski, J. 1959d: Technologia wyrobów żelaznych u dawnych Scytów. Pregla̧d


Techniczny 80/18, 26 - 27.

Piaskowski, J. 1960a: An interesting example of early technology. JISI (London)


194, 336 - 340.

Piaskowski, J. 1960b: Dalsze badania metaloznawcze z okresu póżnohalsztackiego i


wznesnolateńskiego znalezionych na ziemiach Wielkopolski. Fontes Archeol. Posnanienses
11, 94 - 104.
296 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Piaskowski, J.: 1960c: Metaloznawcze badania wyrobów żelaznych z okresu halsz-


tackiego i lateńskiego pochodza̧cych z Malopolski. Materialy Archeol. (Kraków) 9, 197 -
224.

Piaskowski, J. 1960d: Metaloznawcze badania wyrobów żelaznych z okresu halsztack-


iego i lateńskiego znajdzionych na Pomorzu i Mazowszu. Wiadomości Archeol. 1966/3-4,
273 - 286.

Piaskowski, J. 1960e: Metaloznawcze badania wyrobów żelaznych z okresu halsztack-


iego i wzcesnolateńskiego znalezionych na Śla̧sku. Przegla̧d Archeol. 12, 124 - 127.

Paskowski, J. 1960f: Technika gdańskiego hutnictwa i kowalstwa żelaznego X - XIV


w. na podstawie badań metaloznawczych. In: Gdańsk wczesnośredniowieczny vol. II.
Gdańsk.

Piaskowski, J. 1961a: Badania żelaznych wyrobów celtyckich z Karnczy, Sobociska i


Glownina. Silesia Archeol. 3, 80 - 102.

Piaskowski, J. 1961b: Badania metaloznawcze wyrobów żelaznych z cmentarzyska


lultury lużyckiej w Sokolnikach, pow. Tarnobrzeg. Archeologia Polski 6/1, 7 - 19.

Piaskowski, J. 1962a: Metaloznawcze badania dawnych narzȩdzi żelaznych znalezionych


w Krakowie-Nowej Hucie. Wiadomości Archeol. 1962/1, 13 - 20.

Piaskowski, J. 1962b: Metaloznawcze badania wyrobów żelaznych i próbek żużla ze


Śla̧ska Opolskiego z okresu wplywów rzymskich. Przegla̧d Archeol. 15, 134 - 157.

Piaskowski, J. 1962c: Metaloznawcze badania wczesnośredniowiecznych wyrobów


żelaznych i żużla z Sieradza. Prace i Mat. Muzeum Archeol. i Etnograf. w Lodzi 7, 225
- 258.

Piaskowski, J. 1962d: Metaloznawcze badania wyrobów żelaznych z okresu póżnola-


teńskiego i rzymskiego znalezionych na Dolnym Śla̧sku. Silesia Antiqua 4, 198 - 212.

Piaskowski, J. 1962e: Metaloznawcze badania wyrobów żelaznych z osady we Wa̧soszu


Górnym, pow. Klobuck. Wiadomości Archeol. 28/3, 218 - 229.

Piaskowski, K. 1962f: Starożytne kȩsy żelazne z Witowa, pow. Kazimierz Wielka,


w świetle powtórnych badań metaloznawczych. Ancient iron blooms from Witów, distr.
Kazimierz Wielka, in the light of repeated metallographical examinations. Sprawozdania
Archeol. 14, 320 - 329.

Piaskowski, J. 1962g: Technologia i pochodzenie wyrobów żelaznych z pólnocnej


Maloposki i Mazowsza w okresie wplywów rzymskich na podstawie badań metaloznawczych.
Studia z Dziejøw Górnictwa i Hutnictwa 7, 127 - 172.
BIBLIOGRAPHY 297
Piaskowski, J. 1964a: The manufacture of medieval damascened knives, JISI 202,
561 - 568.

Piaskowski, J. 1964b: Metaloznawcze badania narzȩdzi rolniczych z Toporowa, pow.


Wieluń. i Zadowa, pow. Kalisz. Prace i Materialy Muzeum Archeol. i Etnograf. w Lodzi
11, 207 - 232.

Piaskowski, J. 1965a: Correlation between the phosphorus content in ore and slag
and that in bloomery iron. Archaeologia Polona 7, 83 - 103.

Piaskowski, J. 1965b: Niektóre dziwerowane miecze rzymskie na ziemiach Polski. Z


Otchlani Wieków 31/1, 30 - 39.

Piaskowski, J. 1966a: Metaloznawcze badania przedmiotów żelaznych z cmentrazyska


cialopalnego w Zadowicach, pow. Kalisz: Prace i Materialy Muzeum Aecheol. i Etnograf.
w Lodzi 13, 213 - 230.

Piaskowski, J. 1966b: Metaloznawcze badania przedmiotów żelaznych z Mieżan i


Sudaty (LSSR). Wiadomości Archeol. 31/4, 363 - 375.

Piaskowski, J. 1966c: Metaloznawcze badania przedmiotów żelaznych z wczesnośred-


niowiecznej osady w Czela̧dzi Wielkiej, pow. Góra. Silesia Antiqua 8, 150 - 175.

Piaskowski, J. 1966d: Rozwój metaloznawczych badań dawnych przedmiotów żelaz-


nych w Polsce i ich zastosowanie w archeologii. Metallographical investigations of ancient
iron objects in Poland and their application in archaeology. Archeologia Polski X/2, 726
- 750.

Piaskowski, J. 1967a: Metaloznawcze badania starożytnych przedmiotów żelaznych


z dorzeczu Prypeci, Dniestra i Bugu. Materialy Archeol. (Kraków) 8, 197 - 214.

Piaskowski, J. 1967b: Sprawozdanie z metaloznawczych badań starożytnych przed-


miotów żelaznych i żużla z osad w Dalewicach pow. Proszowice i Wólki Lasieckiej, pow.
Lowicz. Sprawozdania Archeol. 18, 356 - 374.

Piaskowski, J. 1969: Metaloznawcze badania starożytnych przedmiotów żelaznych z


Wȩgry, pow. Przasnysz. Étude métallographique des objets en fer de Wȩgra, district de
Przasnysz. Materialy Archeol. (Kraków) X, 73 - 82.

Piaskowski, J. 1970: O produkcji żelaza wysokoniklowego w starożytności. Sur la


production du fer à grande teneur en nickel dans l’antiquité. Acta Archaeol. Carpathica
11/2, 119 - 328.

Piaskowski, J. 1977: Metaloznawcze badania przedmiotów żelaznych z okrsesu hal-


sztackiego z terenu Karpat. Étude métallographique des objets en fer provenant de la
période de Hallstatt dans les Carpathes. Acta Archaeol. Carpathica 17, 259 - 274.
298 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Piaskowski, J. 1980: W sprawie żelaza ‘meteorytowego’ i możliwości jego wytapiania
przes czlowieka. La question du fer ‘météorite’ et les possibilités de sa fonte par l’homme.
Kwartalnik. Historii Kult. Mat. 38/1, 3 - 13.

Piaskowski, J. 1983: Metallographische Untersuchungen zur Eisen- und Stahltech-


nologie in Haithabu. In: Berichte über die Ausgrabungen in Haithabu (Münster) vol. 18,
45 - 63.

Piaskowski, J. 1984: Das Vorkommen von Arsen in antiken und frühmittelalterlichen


Gegenständen. Zeitschrift f. Archäol. (Berlin) 18, 213 - 226.

Piaskowski, J. 1991a: The metallography of the first indirectly smelted iron and steel
found on the territory of Poland, In: Dal basso fuoco all’altoforno. Brescia, 191 - 199.

Piaskowski, J. 1991b: Technologia wczesnośredniowiecznych przedmiotów żelaznych


z wzgórza Wawelskiego. La téchnologie des objets en fer du Haut Moyen age trouvées sur
le colline de Wawel à Cracovie. Studia do dziejów Wawelu Kraków 5, 55 - 92.

Pieta, K. 2000: Ein Depot latènezeitlicher Eisengegenstände aus Liptovská Mara. In:
Metallgwinnung und -verarbeitung in der Antike, 135 - 160.

Pieta, K., Hanuliak, V. 1989: Sklady železných predmetov z konca doby rı́mskej vo
Vyšnom Kubı́ne. Depots von Eisengegenständen aus dem Ende der römischen Kaiserzeit
in Vyšný Kubı́n. AVANS 1987, Nitra 1989, 108 - 109.

Pietzsch, A. 1979: Reitersporen. Arbeits- u. Forschungsberichte z. sächs. Boden-


denkmalpflege 23, 83 - 206.

Pigeau, E. 1994: Ateliers de forge antique et travaux de construction à Chartres


(Eure-et-Loire, site de Barbou). In: La sidérurgie ancienne de l’Est de la France, 137 -
140.

Piggott, S. 1983: The Earliest Wheeled Transport. London.

Pigott, V.C., MacGovern, P. E., Notis, M. R. 1982: The earliest steel from
Transjordan. MASCA Journal 2/2, Archaeometallurgy supplement 1982, 35 - 39.

Piotrovskiy, B. B. 1949: Arkheologiya Zakavkaz’ya. Leningrad.

Pitts, L. F., St.Joseph, J. K. 1985: Inchtuthil: The Roman Legionary Forress.


Britannia Monograph Series 6. London.

Place, V. 1867-1870. Niniveh et l’Assyrie. Paris.

Pleiner, R. 1958: Základy slovanského železářského hutnictvı́ v českých zemı́ch. Die


Grundlagen der slawischen Eisenindustrie in den böhmischen Ländern. Praha.
BIBLIOGRAPHY 299
Pleiner, R. 1961: Slovanské sekerovité hřivny. Die slawischen Axtbarren. Slovenská
archeol. 9/1-2, 401 - 450.

Pleiner, R. 1962: Staré evropské kovářstvı́. Stav metalografického výzkumu. Alteu-


ropäisches Schmiedehandwerk. Stand der metallographischen Forschung. Praha.

Pleiner, R. 1965: Die Eisenverhüttung in der Germania Magna zur römischen Kaiserzeit.
45. Bericht der Röm.-Germ. Kommission 1964 (1965), 11 - 86.

Pleiner, R. 1967a: The Beginnings of the Iron Age in Ancient Persia. Annals of the
Náprstek Museum vol. 6. Prague 1967 [1969], 9 - 79.

Pleiner, R. 1967b: Metallkundliche Untersuchungen der Messerklingen von der früh-


slawischen Siedlung in Dessau-Mosigkau. In: B. Krüger: Dessau-Mosigkau. Berlin, 175 -
189.

Pleiner, R. 1967c: Preliminary evaluation of the 1966 metallurgical investigations in


Iran. In: Investigations at Tal-i-Iblis (J. R. Caldwell ed.). Illinois State Museum Prelimi-
nary Reports vol.9. Springfield. Ill., 340 - 405.

Pleiner, R. 1967d: Die Technologie des Schmiedes in der grossmährischen Kultur.


Slovenská archeol. 15/1, 77 - 188.

Pleiner, R. 1969a: Iron Working in Ancient Greece. Sbornı́k Nár. Tech. Muzea vol.
9. Praha.

Pleiner, R. 1969b: Středověké sı́dliště s kovárnami u Mutějovic. Eine mittelalterliche


Dorfsiedlung mit Schmiedewerkstätten bei Mutějovice, Westböhmen. Památky archeol.
(Praha) 60, 533 - 571.

Pleiner, R. 1969c: Untersuchung eines Kurzschwertes des Luristanischen Typus.


Archäol. Anzeiger (Berlin) 1969/1, 41 - 47.

Pleiner, R. 1970: Zur Schmiedetechnik im römerzeitlichen Bayern. Bayerische Vorge-


schichtsblätter 35, 113 - 141.

Pleiner, R. 1971: The problem of the beginning Iron Age in India. Acta Praehist. et
Archaeol. (Berlin) 2, 5 - 36.

Pleiner, R. 1973: Metallography of Early Artifacts: the Problem of Welding Together


Iron and Steel. In: Antikvariskt Arkiv 53 (Early Medieval Studies 6). Stockholm, 17 -
28.

Pleiner, R. 1976a: Halštatský nůž z Tišic a jeho technologie. Studie Okr. muzea
Praha-východ 1976 6 - 14.

Pleiner, R. 1976b: Nachbildung einer spätkaiserzeitlichen Schwertklinge aus Nor-


300 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
dostböhmen. In: Festschrift Richard Pittioni II (Archaeologia Austriaca Beiheft 4). Wien,
Worms, 130 - 141.

Pleiner, R. 1979a: K vývoji slovanské nožı́řské techniky v Čechách. Zu den Anfängen


der slawischen Messerschmiedekunst in Böhmen. Archeol. rozhledy (Praha) 31, 245 - 256,
354 - 366.

Pleiner, R. 1979b: The Technology of Three Assyrian Iron Artifacts. JNES (Chicago)
38, 83 - 91.

Pleiner, R. 1980a: Early iron metallurgy in Europe. In: The Coming of the Age of
Iron, 375 - 415.

Pleiner, R. 1980b: Zur Technologie der Säbel aus den slawisch-awarischen Gräberfeld-
ern der Südslowakei. In: Rapport du IIIe Congrès d’Archéologie Slave I. Bratislava, 625
- 627.

Pleiner, R. 1981: Die Wege des Eisens nach Europa. In: Frühes Eisen in Europa, 115
- 128.

Pleiner, R. 1982a: Die Herstellungstechnologie der germanischen Eisenwerkzeuge und


Waffen aus den Brandgräberfeldern der Südwestslowakei. Slovenská archeol. 30, 79 - 121.

Pleiner, R. 1982b: Techniky kovářské výroby. Metalografické a chemické rozbory.


Die Schmiedetechnik. In: M. Richter: Hradištko u Davle, městečko ostrovského kláštera.
Praha, 268 - 287, 298 - 300.

Pleiner, R. 1982c: Untersuchungen zur Schmiedetechnik auf den keltischen Oppida.


Památky archeol. (Praha) 73, 86 - 173.

Zur Technik der Messerklingen aus Haithabu. In: Berichte über die Ausgrabun-
gen in Haithabu vol. 18 (Neumünster), 63 - 92.

Pleiner, R. 1986: Über das Eisen der Bronzezeit. In: Veröffentlichungen des Muse-
ums f. Ur- u. Frühgesch. Potsdam 20, 227 - 240.

Pleiner, R. 1991a: Die Technik der Schmiede im mittelalterlichen Prag. Archaeologia


Pragensia (Praha) 11, 239 - 287.

Pleiner, R. 1988 [1991b]: Spätkaiserzeitliche und völkerwanderungszeitliche Stahl-


klingen aus Nordwestböhmen im Lichte der Metallographie. In: Archäometallurgie von
Kupfer un Eisen in Westeuropa, 602 - 610.

Pleiner, R. 1993a: The Celtic Sword. Oxford University Press, Oxford.

Pleiner, R. 1993b: Die Technologie der Messerherstellung in der frühmittelalterlichen


Fürstenburg von Budeč, Böhmen. Památky archeol. (Praha) 84, 69 - 92.
BIBLIOGRAPHY 301

Pleiner, R. 1995: Reflections on the metallography of early artefacts. In: Metallog-


raphy ‘95. Proceedings of the 9th International Conference on Metallography (I. Hrivňák
ed.). Košice, 24 - 29.

Pleiner, R. 1996a: Das frühe Eisen: Von den Kleinwaagemengen zu der ältesten In-
dustrie. EAZ 37, 283 - 291.

Pleiner, R. 1996b: Metallographische Untersuchung des gestempelten römischen


Eisenbarrens E 1170. In: Vitudurum vol. 7. Betráge zum römischen Oberwinterthur.
Ausgrabungen im unteren Bühl. Zürich, Egg, 221 - 228.

Pleiner, R. 1998a: Celtic swords with ‘striped’ damast. In: Metallography ‘98.
Košice, 498 - 500.

Pleiner, R. 1998b: The Technique of Blacksmiths in Noricum and central Europe.


In: Il ferro nelle Alpi. Iron in the Alps. Bienno, 102 - 105.

Pleiner, R. 2000a: Eine Eisenschüssel des sog. Schlesischen Typus aus Grube 95. In:
I. Pleinerová: Die altslwischen Dörfer von Březno bei Louny. Praha, 143 - 146.

Pleiner, R. 2000b: Das Eisen und die Grenze. In: Metallgewinnung und -verarbeitung
in der Antike (Schwerpunkt Eisen), 27 - 31.

Pleiner, R. 2000c: Iron in Archaeology: The European Bloomery Smelters. Praha.

Pleiner, R. 2002a: Metallographische Untersuchung des Schwertes von Bešeňov.


Študijné zvesti Archeol. ústavu SAV, 77 - 82.

Pleiner, R. 2002b: Metalografický výzkum velkomoravské kroužkové zbroje z Břeclavi-


Pohanska. Metallographische Untersuchung von Ringpanzerelementen der grossmährischen
Periode aus Břeclav-Pohansko. Sbornı́k Filozof. fakulty Brněnské univerzity (Brno) M 7,
77 - 81.

Pleiner, R. 2003: Metalografický rozbor železných nástrojů ze Sezimova Ústı́. Me-


tallkundliche Untersuchung an mittlelaterlichen Eisenwerkzeugen von Sezimovo Ústı́. In:
Krajı́c 2003, 173 - 194.

Pleiner, R. 2004: Problems with ancient smithing slag. In: Metallography 2004,
Košice, 664 - 667.

Pleiner, R. sine anno: Metallkundliche Untersuchung an eisernen Waffen und Geräten


aus der spätkeltischen Oppida: Staré Hradisko - Manching - Alesia. Archiv der RGK
Frankfurt/M.

Pleiner, R., Bjorkman, J. K. 1974: The Assyrian Iron Age: The History of Iron in
Assyrian Civilization. Proceedings of American Phil. Society 118/4, 283 - 313.
302 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pleiner, R., Fluzin, P., Mangin, M., Billon, M., Dillmann, P., Véga, P.,
Rabeisen, É. 2003: Lingots et couteaux en fer f’Alésia: études archéometriques de
pièces inédits. Revue Archéologique de l’Est 52, 91 - 130.

Pleiner, R., Huml, V. 1993: Punchmarks on High Medieval iron artefacts. In:
Bloomery Ironmaking During 2000 Years III, 79 - 89.

Pleiner, R., Pelikán, J., Bartuška, M. 1971. Untersuchung einer Eisenschlacke


aus Haithabu. In: Berichte über die Ausgrabungen in Haithabu vol. 5 (K. Schietzel ed.).
Münster, 110 - 112.

Pleiner, R., Plzák, F., Quadrat, O. 1956: Poznámky k výrobnı́ technice staroslo-
vanských čepelı́. Bemerkungen zur Erzeugungstechnik altslawischer Klingen. Památky
archeol. (Praha) 47/2, 314 - 334.

Pleiner R., Princ, M. 1984: Zur latènezeitlichen Eisenverhüttung und die Unter-
suchung einer Rennschmelze in Mšec, Böhmen. Památky archeol. (Praha) 75, 133 - 180.

Pletneva, S. A. 1996: Metallurgicheskiy i kuznechnyy komplexy Sarkela. The met-


allurgical and blacksmith’s complexes of Sarkel. Rossiyskaya arkheolgia 2/1996, 182 - 197.

Polesskikh. A. M. 1961: Nachodka skifskogo metcha v Penzenskoy oblasti. Sovet.


arkheol. 1/1961, 257 - 259.

Polla, B. 1962: Stredoveká zaniknutá osada na Spiši (Zalužany). Bratislava.

Popović, I. 1988: Antičko orude v Srbiji. Les outils en fer de Serbie. Beograd.

Prehistoric and Medieval Iron Smelting in Sweden. Aspects of Technology and Science
(L. Chr. Nørbach ed.). Aarhus University Press, Aarhus 2003.

Presslinger, H., Köstler, H. J. 1991: Der Werkstoff Stahl im Altertum. Ferrum -


Nachrichten aus der Eisenbibliothek +GF+ Stiftung. Schaffhausen 1991/3, 18 - 26.

Princ, M. 1981: Dı́lna kováře v Hradišti u Českých Lhotic. Schmiede auf dem Op-
pidum Hradiště bei České Lhotice. In: Praehistorica VIII, Varia Archaeologica 2. In
honorem J. Filip. Praha, 209 - 215.

Prostřednı́k, J., Hošek, J. 2001: A smithy workshop at the Castle Trosky. In:
Archaeometallurgy in The Central Europe III, Acta Metallurgica Slovaca 7, Spec. Issue
2/2001, 27 - 35.

Pulsifer, H. B. 1931: Low-carbon steel. Transactions of the American Institute of


Mining and Metallurgical Engineers, Iron and Steel Division (New York), 31 - 35.
BIBLIOGRAPHY 303
Pusch, R. 1976: Geschichte der Metallographie. Düsseldorf.

Py, M. 1978: L’oppidum des Castels à Nages (Gard). Fouilles 1957 - 1974. XXXVe
supplement à Gallia. Paris.

Rädeker, W., Naumann, F. K. 1961: Untersuchung vor- und frühgeschichtlichen


Spitzbarren. Archiv f. f. Eisenhüttenwesen (Düsseldorf) 32/9, 587 - 595.

Rageth, J. 1982: Die römischen Schmiedegruben von Riom GR. Archäologie der
Schweiz 5, 202 - 208.

Rasl, Z. 1988: Záěsné zámky dvou tisı́ciletı́ z hlediska technického vývoje. Padlocks
and their technical evolution during two thousend years. Národnı́ tech. muzeum 1978-
1988 2. Studie z dějin techniky, 153 - 225, 324.

Rasl, Z. 1995: Nástin vývoje zámkových mechanismů. In: Černé řemeslo v průběhu
staletı́. Národ. tech. muzeum Praha, 4 - 36.

Raspopova, V. I. 1980: Metallicheskiye izdeliya rannesrednevekogo Sogda. Leningrad.

Rauhut, L. 1956: Sprawozdanie z badań wczesnośredniowiecznego ośrodka metalur-


giczno-kowalskiego we wsi Kamionka Nadbuźna, pow. Ostrów Mazowiecki. An early
metallurgical centre at Kamionka Nadbużna, Ostrów Mazowiecki district. Wiadomości
Archeol.23, 342 - 352.

Rauhut, L. 1957: Studia i materialy do historii starożytnego hutnictwa żelaza w


Polsce. Studia z Dziejów Górnictwa i Hutnictwa 1-1, 183 - 293.

Rebière, J., Rémy, P., Guillot, I., Benoit, P. 1995: Les enclumes tas. Le cas de
Jouars-Ponchartrain. In: Paléometallurgie du fer & Cultures, 501 - 508.

Rech, M. 1980: Eine Villa rustica im Hambacher Forst, Kr. Düren. Bonner Jahrb.
180, 461 - 491.

Recherches sur l’économie du fer en Méditerranée nord-occidentale. Monographies in-


strumentum 4, Éd. Mon. Margoil, Montagnac.

Rehder, T. 1981: Ancient Carburization of Iron and Steel. Archaeometry 3/1, 27 - 37.

Rehren, Th., Hauptmann, A. 1994: Römische Eisenblöcke von der Saalburg. Un-
tersuchungen zur Fertigungstechnik. Saalburg Jahrbuch 47, 79 - 85.

Reichmann, Chr. 1984: Eine mittelalterliche Schmiede aus Bocholter Kirchhof. Aus-
grabungen u. Funde in Westphalen-Lippe 2, 69 - 100.

Reinerth, H. 1931: Ein Depotfund der Spätlatènezeit von Dürnau. Nachrichtenblatt


f. deutsche Frühzeit 7. 63 sq.
304 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Rétif, M. 1998: Indices de métallurgie protohistorique du fer à Martigues (Bouche du


Rhône). In: Recherches sur léconomie du fer, 96 - 109.

Richardson, T., Starley, D. 2002: The helmet from Stratto Billage, Bedfordshire.
Royal Armouries Yearbook 7, 15 - 21.

Richter, K. 1997: Die Metalle im Alten Orient unter besonderer Berücksichtigung


altbabylonischen Quellen. Ugarit Verlag, München.

Ridgway, D. 1973: Metalworking at Pithekoussai, Ischia (NA), Italy. Archeol. rozh-


ledy (Praha)25, 456.

Rieth, A. 1942: Eisentechnik der Hallstattzeit. Leipzig.

Roesdahl, E. 1977: Fyrkat. En jysk vikingeborg II. Oldsagerna og gravpladsen.


Fyrkat. A Viking fortress in Jutland II. The finds and cemetery. København.

Rosetti, D. V. 1960: Un deposit de unelte, cı̂teva stampile anepigrafice si o monadă


din e doua epocă a fierului, SCIV 11/2, 391 - 403.

Rostoker, W. 1987: White cast iron as a weapon and tool material. Archeomaterials
1/2, 145 - 148.

Rovira, C. 1998: Le travail du fer en Catalogne du VIIe au Ier s. avant notre ère. In:
Recherches sur la économie du fer, 65 - 75.

Rozanova, L. S., Terekhova, N. N. 2002: K probleme kavkazskikh i mestnykh tradi-


ciy v tekhnologii izgotovleniya zheleznykh izdeliy iz starshego akhmylovskogo mogilnika.
On the problem of Caucasian and local traditions in the technology of manufacturing of
iron objects from starshee Akhmylovo cemetery. KS 213, 72 - 80.

Rozanova, L. S., Terekhova, N. N. 2003: Iron-working in classical sites of the


North Pontic area (according to the materials from Gorgippia). In: Archaeometallurgy
in Europe, Milan, 63 - 70.

Różański, W. 1958: Badania przedmiotów metalowych z grobów cialopalnych. Stu-


dia z Dziejów Górnictwa i Hutnictwa 2, 99 - 114.

Rump, P. 1964: Herstellung westfälischer Zieheisen. Ein Beitrag zur Geschichte des
Drahtziehens. Stahl u. Eisen (Düsseldorf) 84, 1260 - 1265.

Rupe, H., Müller, F. 1916: Chemische und metallographische Untersuchungen


prähistorischer Eisenfunde. Verhandlungen der Naturforsch. Gesellschaft Basel, 108 sq.

Rusev, R. 1961: Verkhu tekhnologiyata na nyakoye zhelezni predmeti ve XII - XIII


v. Arkheologiya (Sofia) 3/2, 6 - 14.
BIBLIOGRAPHY 305

Rustoiu, A. 2000: Outils en fer pour le travail les métaux non ferreux in Dacie
préromaine (Ier siècle avant J.-C. - Ier siècle ap. J.-C.). In: Iron, Blacksmiths and Tools,
233 - 241.

Rybakov, B. A. 1948: Remeslo drevney Rusi. Moskva.

Rybová, A., Motyková K. 1983: Der Eisendepotfund der Latènezeit von Kolı́n.
Památky archeol. (Praha) 74, 96 - 174.

Sacken, E. von 1868: Das Grabfeld von Hallstatt. Wien.

Sal’dau, P. Ya., Gushtina, A. F. 1932: Primenenye metallografii v arkheologii.


Soobchtcheniya GAIMK (Moskva) No 3/4, 49 sq.

Salin, E. 1956: Les techniques métallurgiques après les grandes invasions. In: Le fer
à travers les âges. Nancy, 45 - 55.

Sandars, N. K. 1971: From Bronze to Iron. A Sequel to a Sequel. In: The European
Community in Later Prehistory (Studies in honour of C.F.C. Hawkes). London, 1 - 29.

Sankot, P. 2003: Les épées du début de La Tène en Bôheme. Fontes Archaeologici


Pragenses (Praha), vol. 18.

Sankot, P., Vojtěchovská, I. 1984: An early La Tène hoard of iron artifacts at


Chýnov near Prague. Archeol. rozhledy (Praha) 34, 683.

Saska, L. F., Groh, F. 1949: Mythologie Řeků a Řı́manů. Praha (9th edition).

Schaaber, O. 1963: Beiträge zur Frage des norischen Eisens. Metallkundliche Grund-
lagen und Untersuchungen an Funden vom Magdalensberg. Carinthia I (Klagenfurt) 153,
129 - 208.

Schaaber, O. 1964: Metallkundliche Untersuchung an Fundstücken vom Magdalens-


berg. Archiv f. d. Wisenhüttenwesen (Düsseldorf) 35/6, 502 - 506.

Schäfer, K. 1984: Ein spätlatènezeitlicher Eisenbarrendepot aus Säffig, Kreis Mayen-


Koblenz. Archäol. Korrespondenzblatt 14/2, 163 - 168.

Schaeffer, C. F. A. 1948: Stratigraphie comparée et chronologie de l’Asie occidentale


(IIIe et IIe millénnium). Paris.

Schaeffer, F. A. 1927: Un dépôt d’outils et un trésor de bronze de l’époque Gallo-


Romain décpuvert à Selby (Bas Rhin). Publications du Musée de Hagenau (Alsace).
Hagenau.

Schmid, E. D. 2003: Link Details from Articles of Mail in the Wallace Collection.
306 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
The Journal of the Mail Research Society (St. Cloud, USA) I/1 , 2 - 20.

Schmid, W. 1932: Die römische Poststation Noreia in Einöd. Jahreshefte des Österreich.
Archäol. Inst. 27, 194 - 222.

Schmid, W. 1937: Ulaka. Japodska naselbina nad Starym Trgem pri Ložu. Glasnik
muzejskog društva za Slovenija (Ljubljana) 18, 17 - 32.

Schmitz, F. 1930: Orientalischer Damaststahl. Beiträge zur Geschichte der Technik


und Industrie, Jahrbuch des Vereines deutsch. Ingenieure (Berlin) 20. 81 - 86.

Schoknecht, T. U. 1977: Menzlin. Ein frühgeschichtlicher Handelsplatz an der Peene.


Berlin.

Schönberger, H. 1967: Ein Eisendepot, römische Flosfesseln und andere Funde des
Kastells Heilbronn-Böckingen. Fundberichte aus Schwaben N.F. 18/I, 131 - 157.

Schönberger, H. 1978: Kastell Oberstimm. Die Grabungen von 1968 bis 1971.
Berlin. C. Böhne, Metallurgie, pp. 164 227.

Schoppa, H. 1964: Eine römische Strassenstation bei Kriftel, Maintaunuskreis. Fund-


berichte aus Hessen 4, 98 - 116.

Schreyer, S., Graf, M. 1995: Rheinau ZH. Eine keltische Schmiedewerkstatt und
andere Siedlungsreste aus keltischer Zeit an der Austrasse. Archäologie d. Schweiz 18/1,
33.

Schuldt, E. 1980: Handwerk und Gewerbe des 8. bis 12. Jahrhunderts in Mecklen-
burg. Schwerin.

Schulz, E. H. 1955: Über den Werkstoff des Schweisseisenzeitalters. Archiv f. d.


Eisenhüttenwesen (Düsseldorf) 26, 265 - 371.

Schulz, E. H. 1959: Über die metallkundliche Untersuchung einiger römischen Schwert-


klingen. Technische Beiträge zur Archäologie I. Mainz, 46 - 64.

Schulz, E. H., Hülsbruck, W. 1927: Über die Randentkohlung der Kohlenstoffstähle.


Archiv f. d. Eisenhüttenwesen (Düsseldorf) I, 225 - 240.

Schürmann, E. 1959: Untersuchungen an Nydam Schwertern. Archiv f. d. Eisen-


hüttenwesen (Düsseldorf) 30, 121 - 126.

Schuzany, C. 1986: Der römische Gutshof von Biberist-Spitalhof. Ein Vorbericht.


Jahrbuch d. Schweiz. Gesellaschaft f. Ur- u. Frühgesch. 69, 199 - 220.

Schuzany, C. 1994: Les forges de la villa rustica de Biberist-Spitalhof (Suisse). In:


La sidérurgie ancienne de l’Est de la France, 143 - 157.
BIBLIOGRAPHY 307

Schwab, R. 2000: Überlegungen zur Eisenversorgung des Oppidums von Manch-


ing basierend auf metallkundlichen Untersuchungen von Waffen und Geräten. Berliner
Beiträge zur Archäometrie 17, 5 - 44.

Schwab, R. 2002: Evidence for carburized steel and quench-hardening in the ‘Celtic’
oppidum of Manching. HM 36/1, 6 - 16.

Scott, B. G. [1990]: Early Irish Ironworking. Ulster Museum, Belfast.

Sedov, V. V. 1960: Sel’skiye poseleniya central’nykh rayonov Smolenskoy zemli (VIII


- XV vv.). MIA vol. 92. Moskva, 74 sq.

Seek, O. 1876 (1962): Notitia Dignitatum. 2nd. edition. Minerva, Frankfurt/M 1962.

Selirand, J., Tönisson, E. 1984: Through Past Millennia. Archaeological Discover-


ies in Estonia. Tallin.

Senn-Luder, M. 1998: Schlacken und Schmelzgefässe als Spiegel der Metallgewerben


im alten Zug (cooperation with V. Serneels, contributions by A. Rehazek, J. Schibler, M.
Veszeli). Tugium (Zug) 14, 113 - 154.

Serneels, V. 1993: Archéometrie des scories de fer: Recherches sur la sidérurgie en


Suisse occidentale, Cahiers d’archólogie romade vol. 61. Lausanne.

Serneels, V. 1996: see T. Geiger.

Serneels, V. 1998: Les ateliers sidérurgiques d’Autun-Lycée Militaire. In: Les métaux
antiques, 23 - 27.

Serneels, V., Mauvilly, M. 2001: The Early La Tène Metallurgical Workshop at


Sévaz FR/Switzerland: An attempt of quantification. In: Archaeometallurgy in The Cen-
tral Europe III. Acta Metallurgica Slovaca vol. 7, 268 - 277.

Serning, I. 1965: Vor- und frühgeschtliches Eisengewerbe im schwedischen Järnbäraland.


In: Vita pro Ferro (Festschrift Durrer, W. U. Guyan ed.). Schaffhausen, 73 - 90.

Serning, I. 1973: Förhistorisk järnhantering i Dalarna. Jernkontorets Forskning Ser.


H9. Stockholm.

Seyffert, O. 1902: A dictionary of classical antiquities, mythology, religion, literature


& art (with contribution by H. Nettleship and J. E. Sandys). London, New York.

Shramko, B. A. 1969: Orudiye skifskoy epochi dlya obrabotki zheleza. Les outils de
fer de l’époque scythique pour le travail du fer. Sov. arkh. 1963/3, 53 - 70.

Shramko, B. A. 1981: Die ältesten Eisenfundstücke im Osteuropa. In: Das frühe


308 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Eisen, 109 - 114.

Shramko, B. A. 1987: Belskoye gorodichtche skifskoy epochi (gorod Gelona). Kyiv.

Shramko, B. A., Fomin, L. D., Solncev, L. A. 1965: Pervya nachodka izdeliz iz


meteoritnogo zheleza v vostochnoy Evrope. La première trouvaille d’un produit de fer de
météorite en Europe de l’Est. Sov. arkheol. 1965/4, 57 - 74.

Shramko, B. A., Fomin, L. D., Solncev, L. A. 1971: Novyye issledovaniya


tekhniki obrabotki zheleza v Skifii. Nouvelles recherches concernant la technique du tra-
vail du fer en Scythie. Sov. arkh. 1971/4, 140 - 152.

Shramko, B. A., Petrichenko, O. M., Solncev, L. A., Fomin, L. D. 1961:


Doslizhdeniya starorus’kikh zaliznykh vyrobiv Doneckogo horodichtcha. Narysy ist. tekhn.
(Kyiv) 7, 74 - 87.

Shramko, B. A., Solncev, L. A. 1991: Metallograficheskiye issledovaniya zheleznykh


izdeliy iz skifskogo kurgana Chertomlyk. In: A. Yu. Alekssev, B. Yu. Murzin, R. Rolle:
Chertomlyk (Skifskiy carskiy kurgan IV veka do n. e.). Kyiv, 390 - 397.

Shramko, B. A., Solncev, L. A., Fomin, L. D. 1963: Tekhnika obrabotki zheleza


v lesostepnoy i stepnoy Skifii. Sov. arkh. 1963/4, 36 - 57.

Shramko, B. A., Solncev, L. A., Stepanskaya, R. B., Fomin, L. J. 1974: K


voprosu o tekhnike izgotovleniya sarmatskikh mechey i kinzhalov. A propos de la tech-
mique de fabrication des épées et poignards Sarmates. Sov. arkh. 1974/1, 181 - 190.

La sidérurgie ancienne de l’Est de la France dans son contexte européen. Archéologie


et archéometrie (M. Mangin ed.) Paris 1994.

Siegelová, J. 1984: Gewinnung und Verarbeitung von Eisen im Hettitischen Reich im


2. Jahrtausend v. u. Z. Annals of the Náprstek Museum (Prague), 71 - 168.

Sievers, S. 1992: Die Kleinfunde. In: Ergebnisse de Ausgrabungen 1984 - 1987 in


Manching (F. Meier ed.). Stuttgart, 136 - 213.

Sievers, S. 1999: Vortrag zur Jahressitzung der Römisch-Germanischen Kommission:


Manching - Aufstieg und Niedergang einer Keltenstadt. Berichte der RGK 80, 7 - 23.

Sievers, S. 2000: Vorbericht über die Ausgrabungen 1998 - 1999 im Oppidum von
Manching (contributions by R. Gebhard, M. Leicht, R. Schwab, j. Völker, B. Weber, B.
Ziegaus). Germania 78/2, 355 - 394.

Sim, D. 1992: Some misunderstandings concerning the blacksmith’s tongs. HM 26,


63 - 65.

Sim, D., Ridge, I. M. L. 2000: Examination of a moulding plane from Vindolanda.


BIBLIOGRAPHY 309
HM 34/2, 77 - 82.

Simoni, K. 1982: Skupni nalaz oruda i oružja iz Nartskih Novaka. Ein Sammelfund
von Waffen und Werkzeugen aus Nartski Novaki, Vjesnik Arheološkog Muzeja v Zagrebu
3/15, 251 - 266.

Skurczyński, S. 1958: Szaniec, pow. Busko (Nowe odkrycia). Z Otchlani Wieków


24, 43 - 44.

Die Slawen in Deutschland. Geschichte der slawischen Stämme westlich von Oder und
Neisse vom 6. bis 12. Jh. (J. Herrmann ed.). Berlin.

Slivka, M. 1978: Stredoveké hutnı́ctvo a kováčstvo na Slovensku 1. Historica Carpat-


ica (Košice), ser. C. IX, 217 - 263.

Slivka, M. 1980: Stredoveké kováčstvo i hutnı́ctvo na Slovensku 2, Historica Carpat-


ica (Košice) 11, ser. C., 218 - 288.

Smith, C. S. 1960: A History of Metallography. The Development of Ideas on the


Structures of Metals. Univ. of Chicago Press. Chicago (reprinted 1988).

Smith, R. 1905: The ancient British iron currency. Proceedings of the Society of
Antiquaries (London) 20, 179 - 195.

Smith, R. H., Maddin, R., Muhly, J. D., Stech, T. 1984: Bronze Age steel from
Pella, Jordan. Current Anthropology 25, 234 - 236.

Snodgrass, A. M. 1971: Dark Age of Greece. An Archaeological Survey of the


Eleventh to the Eight Centuries BC. Edinburgh University Press. Edinburgh.

Snodgrass, A. M. 1980: Iron and Early Metallurgy in the Mediterranean. In: The
Coming of Iron, 335 - 374.

Sönnecken, M. 1971: Bericht über die Ausgrabung einer Schmiede des 12. und 15.
Jahrhunderts bei Haus Rhade, Stadt Kierspe, Kreis Lüdenscheid. Der Märker 20/1, 19 -
21.

Sönnecken, M. 1977: Forschungen zur spätmittelalterlich-neuzeitlichen Eisenherstel-


lung in Kierspe, Märkischer Kreis. Fachausschussbericht 9006 VDEh. Düsseldorf.

Sorokin, S. 1957: Novyy arkheologicheskiy material po istorii russkogo zhelezode-


latelnogo proizvodstva. Soobchenniya Gos. ermitazha 11, 22 - 24.

Souchopová, V. 1995: Počátky západoslovanského hutnictvı́ železa ve světle pra-


menů z Moravy. The Beginnings of the metallurgy of Iron in the Light of the Sources
from Moravia. Studie Archeol. ústavu AV ČR v Brně vol. XV/1. Brno.
310 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Souchopová, V. 1996: Nové náhledy na objev kovárny v jeskyni Býčı́ skála. Neue
Ansichten über die Entdeckung einer Schmiede in der Höhle Býčı́ skála. In: Z dějin hut-
nictvı́ vol. 25 (Rozpravy Nár. tech. muz. v Praze 142). Praha, 6 - 8.

Souvenir: National Seminar on Indian Archaeometallurgy, 2 - 4 Oct. 1991. Varanasi


1991.

Sperl, G. 1980: Über die Typologie urzeitlicher, frühgeschichtlichen und mittelalter-


lichen Eisenhüttenschlacken. Studien zur Industrie-Archäologie vol. VII. Österr. Akademie
d. Wiss., Wien.

Stankus, J. 1974: Geležnies dirbniu gamybos raida Lietuvoje. Lietuvos Istorios me-
trastis 1973. Vilnius 1974, 5sq.

Stech, T., Muhly, J. D., Maddin, R. 1985: The Analysis of Iron Artifacts from
Palaeopaphos-Skales. Report of the Dep. of Antiquities, Cyprus, 192 - 202.

Stech-Wheeler, T., Muhly, J. D., Maxwell-Hyslop, K. R., Maddin, R. 1981:


Iron at Taanach and Early Iron Metallurgy in the Eastern Mediterranean. AJA 85, 245
- 268.

Stjernquist, B. 1961: Simris II. Bronze Age Problems in the Light of Simris Excava-
tion. Bonn, London.

Stoumann, I. Vikinga Landsbyen i Saedding. Marl og Montre, 30 - 42.

Stoumann, I. 1979: A Viking Age village near Esbjerg. Acta Archaeol. Copenhagen
50, 95 - 118.

Straube, H. 1996: Ferrum Noricum und die Stadt auf dem Magdalensberg. With
contributions by H. Dolenz and G. Piccotini. Wien, New York.

Sulowska, M. 1977: Klopoty se skarbem. Z Otchlani Wieków 43, 1977/1, 32 - 38.

Svane, H. 1991: Rhombic iron axes and axiform currency bars from Norway. In:
From Bloom to Knife, 31 - 33.

Svobodova, J. 1977: Im Busch und Savannenland der Fulbe von Senegal. Beiträge
aus dem Staatlichen Museum f. Völkerkunde Dresden 1977, 31 - 57.

Swiss, A., McDonnell, J. G. 2004: Evidence and interpretation of cold working in


ferritic iron. In: Archaeometallurgy in Europe, Milan, 209 - 217.

Swoboda, R. M. 1986: Die spätrömische Befestigung Sponeck am Kaiserstuhl. München.

Szabó, N. 1966: A keltöspirimas alaká vasrudak Kérdéséher. Sur la question des bar-
res de fer en forme de double pyramides (Doppelpyramidenbarren). Archeologiai Értesitö
BIBLIOGRAPHY 311
93/2, 249 - 258.

Székély, Z. 1966: Beiträge zur Kenntnis der Frühhallstattzeit und zum Gebrauch des
Eisens in Rumänien. Dacia N. S. 10, 209 - 219.

Taffanel, o. and E. 1967: Les épées à sphères du Cayla à Mailhac (Aude). Gallia
25/1, 1 - 10.

Tauber, J. 1992: Zum Stand der Eisenarchäologie im Kanton Basel-Landschaft. Mine-


ria Helvetica 1992, 22 - 30.

Taus, M. 1963: Ein spätlatènezeitliches Schmiedegrab aus St. Georgen am Steinfeld,


p. B, St. Pölten. Archaeol. Austriaca 34, 13 - 16.

Teodor, S. 1979: Dépot d’outils d’époque La Tène de Lozna, dép. de Botoşani. In-
ventaria Archaeologica, Roumania vol. 11. Bucureşti.

Teodor, S. 1980: Das Werkzeugdepot von Lozna (Kr. Botoşani). Dacia N.S. 24, 133
- 150.

Terekhova, N. N., Rozanova, L. S., Zav’yalov, V. I., Tolmacheva, M. M.


1997: Ocherki po istorii drevney zhelezoobrabotki v Vostochnoy Evrope. Essays on His-
tory of Ancient Ironworking in Eastern Europe. Metallurgiya, Moskva.

Thålin, L. 1966: Redogörelse för tekniska undersögningar. In: Helgö och Metallun-
dersökningarna. Årsrapport 1966. Stockholm, 46 - 80.

Thålin-Bergman, L. 1979: Blacksmithing in Prehistoric Sweden. In: Iron and Man


in Prehistoric Sweden, 99 - 133.

Thålin-Bergman, L. 1983: Der wikingerzeitliche Werkzeugkasten vom Mästermyr


auf Gotland. In: Das Handwerk in vor- und frühgeschichtlicher Zeit II (H. Jankuhn et al.
eds). Göttingen, 193sq.

Thiollier, F and N. 1899: Fouilles du Mont Beuvray (ancienne Bibracte). Album.


St. Étienne.

Thomsen, R. 1964: Forsøg på rekonstruktion fortidiger smedeprocesser. Attempt at


reconstruction of prehistoric forging processes. Kuml 1964, 62 - 85.

Thomsen, R. 1971a: Essestein und Eisenschlacken aus Haithabu. In: Berichte über
die Ausgrabungen in Haithabu vol. 5, Untersuchungen zur Technologie des Eisens (L.
Schietzel ed.). Münster, 100 - 109.

Thomsen, R. 1971b: Metallographische Untersuchung an drei wikingerzeitlichen


Eisenäxten aus Haithabu. Ibidem 30 - 57.
312 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Thomsen, R. 1971c: Metallographische Untersuchung an wikingerzeitlichen Eisen-
barren aus Haithabu.Ibidem, 9 - 29.

Thomsen, R. 1971d: Metallographische Untersuchung einer wikingerzeitlichen Lanzen-


spitze aus Haithabu. Ibidem, 58 - 83.

Thomsen, R. 1992: Metallografiska undersøgelser af svaerd og spydspidser fra mose-


funde i Illerup og Nydam. Patternwelded swords from Illerup and Nydam. Aarsbøger
1992, 281 - 310.

Timpel, W. 1963: Ergebnisse einer Neuuntersuchung des karolingischen Schwertes


von Ottmannshausen. Ausgrabungen u. Funde 8/5, 262 - 267.

Tizzoni Cucini, C. 1999: Ponte di Val Gabbia III: la forgia e i bassofuochi fra Tar-
doantico e alto Medioevo. In: La miniera perduta. Bienno, 93 - 139.

Točı́k, A. 1983: Vel’komoravský železný depot z Čeboviec. Grossmährischer Eisende-


potfund von Čebovce. Štud. zvesti Archeol ústavu SAV 20, 207 - 230.

Traditions and Innovations in the Early Medieval Ironproduction. Papers of the So-
pron and Somogyfajsz Conference (J. Gömöri ed.). Sopron, Somogyfajsz 1999.

Trňáčková, Z. 1985: Ein Hortfund von Metallgegenständen aus der späteren Kaiserzeit
und frühen Völkerwanderungszeit aus Mušov. Památky archeol. (Praha) 76/2, 278 - 285.

Tschumi, O. 1953: Urgeschichte des Kantons Bern (Alter Kanton Steil). Einführung
und Fundliste. Bern, Stuttgart.

Tylecote, R. F. 1962: Metallurgy in Archaeology. London.

Tylecote, R. F. 1976L A History of Metallurgy. London.

Tylecote, R. F. 1986: The Prehistory of Metallurgy in the British Isles. The Inst. of
Metals, London.

Tylecote, R. F. 1987: The early history of metallurgy in Europe. London, New York.

Tylecote, R. F., Gilmour, B. 1986: The Metallography of Edge-Tools and Edged


Weapons. BAR British Series vol. 155. London.

Tylecote, R. F., Thomsen, R. 1973: The Segregation and Surface-Enrichment of


Arsenic and Phosphorus in Early Iron Artifacts. Archaeometry 15, 193 - 198.

Uenze, S. 1992: Sadovec. Die spätantiken Befestigungen von Sadovec I - II. München.

Ulbert, G. 1984: Cáceres el Viejo - ein spätrepublikanischer Legionslager in Spanisch–


Extremadura. Mainz am Rhein.
BIBLIOGRAPHY 313

Ulbert, G. 1988: Die frühkaiserzeitliche Siedlung auf dem Auerberg. In: H.-J.
Küster: Vom Werden einer Kulturlandschaft. Vegetationsgeschichtliche Studien am Auer-
berg, Südbayern. Quellen und Forschungen zur prähistorischen und provinzialrömischen
Archäologie. Weinheim.

Un quartier antique d’artisanat métallurgique à Autun. La site du Lycée militaire. P.


Chardron-Picault, M Pernot eds. dAf vol. 76. Paris.

Unger, J. 1999: Život na lelekovickém hradě ve 14. stoletı́. Antropologická sociokul-


turnı́ studie. Das Leben auf der Burg Lelekovice im 14. Jahrhundert: Anthropologische
soziokulturelle Studie. Brno.

Unglik, H. 1991: Observations on the structures and formation of microscopic smithing


residues from Bixby Blacksmith Shop at Barre Four Corners, Massachussets, 1824-55. HM
25/2, 92 - 98.

Uran, L. 1983: Contribution à l’étude de la paléometallurgie du fer: Structures d’épées


celtiques. Thesis. Université de technologie. Compiègne.

Urban, O. H. 2000: Ein mittel/junglatènezetlicher Eisendepotfund von Falkenstein,


NÖ. In: Metallgewinnung und -verarbeitung in der Antike, 195 - 199.

Ustohal, V., Vı́t, J., Ptáčková, M. 2001: The band-like structure in iron articles
from archaeological sites. In: Archaeometallurgy in The Central Europe III, 209 - 213.

Valter, I. 1981: Árpád-kori kovácsmühely Csatáron. Arpád period forge at Csatár.


In: Iparrégészet I. Veszptém, 123 - 131.

Van Beek, G. W. 1993: Jemmeh, Tell. In: The New Encyclopaedia of Archaeolog-
ical Excavations in the Holy Land. A Paramount Communication Comp. New. York,
London. Toronto, Sydney, Tokyo, Singapore, 667 - 674.

Varoufakis, G. J. 1970: Examination of three steel spear tips of the 7th and 6th
cent. BC. Archiv f.d. Eisenhüttenwesen (Düsseldorf) 41, 1023 - 1026.

Varoufakis, G. J. 1992: The iron clamps and dowels from the Parthenon and
Erechthion. HM 26, 1 - 18.

Važarova, Z. N. 1960: Železarskata rabotilnica u ”Selišča”-Preslav. Slavia Antiqua


7, 398 - 405.

Vercingétorix at Alésia. Paris 1994.

Verna, C. 1995: Les mines et les forges des Cisterciennes en Champagne méridionale
et en Bourgogne du Nord, XIIe - XVe siècles. Vulcain. Paris.
314 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Vetters, H. 1966: Ferrum Noricum. Anzeiger d. österr. Akad. d. Wiss., phil.-hist.
Klasse 167sq.

Vidal, L. 1998: Un depotoir de rejets de forge à Nı̂mes (Gard), rue Condé. In:
Recherches sur l’économie du fer, 153 - 154.

Vierck, H. 1983: Ein Schmiedeplatz aus Alt-Ladoga und der präurbane Handel zur
Ostsee vor der Wikingerzeit. Münsterische Beiträge zur antiken Handelsgeschichte II/1,
3 - 67.

Vitljanov, S. 1990: Das Schmiedehandwerk in Bulgarien von VII.- XIV. Jh. Slavia
Antiqua 32 1989/1990, 145 - 170.

Vorbericht über die Ausgrabungen 1996-1997 im Oppidum von Manching. S. Sievers


et al.. Germania 76/2, 619 - 672.

Voss, O. 1993: Snorup. Et jernudvindingsområde i Sydvestjylland. Snorup - an iron


smelting area in south-west Jutland. Nationalmuseet Arbeidsmark 1993, 97 - 111.

Voss, O. 1996: An Iron Age Settlement with Iron Production of the 4th to the 6th cen-
tury AD. In: Early Iron. Netvork for tidlig jernteknologi (H. Lyngström ed.). København,
19 - 28.

Vouga, P. 1925: La Tène. Monographie de la station publiée au nom de la commission


des fouilles de La Tène. Leipzig.

Voznesenskaya, G. A. 1965a: Metall Troyickogo gorodichtcha. Arkheologiya i es-


testvennyye nauki. Moskva, 123 - 138.

Voznesenskaya, G. A. 1965b: Stal’nyye nozhi drevnego Lyubetcha (metallografich-


eskye issledovaniya). KS (Moskva) 104, 145 - 148.

Voznesenskaya, G. A. 1999: The technology of the blacksmith’s production in the


rural settlements of Southern Russia. In: Archaometallurgy in The Central Europe,
Košice, 108 - 114.

Voznesens’ka, G. A., Nedopako, L. T., Pan’kov, S. V. 1996: Tchorna me-


tallurhiya ta metalloobrobka schodnoevropeys’koho lisostepy za dobi rannikh Slov’yan i
kyievskoy Rusi (druga polovina I tis. - persha tchvert’ II tis). Kyiv.

Wacher, J. 1974: The Towns of Roman Britain. London.

Wagner, D. B. 1997: The earliest use of iron in China. In: Metals in Antiquity (S.
H. M. Young, P. Budd eds.), preprint.

Walander, A. 1975: Ett ämnesjärnsfynd frå n Eketorp borg på Öland. A Find of
Iron Blanks from Eketorp Fort, Öland. Fornvännen (Stockholm) 70, 147 - 155.
BIBLIOGRAPHY 315

Waldbaum, J. C. 1978: From Bronze to Iron. The Transition from the Bronze Age
to the Iron Age in the Eastern Mediterranean. Göteborg.

Waldbaum, J. C. 1980: The First Archaeological Appearance of Iron and the Tran-
sition to the Iron Age. In: The Coming of Iron, 69 - 98.

Waldbaum, J. C. 1999: The Coming of Iron in the Eastern Mediterranean. MASCA


Research Paper vol. 16 (V. C. Piggot ed.). The University Museum of Pennsylvania, 27
- 57.

Walke, N. 1965: Das römische Donaukastell Straubing - Sorviodurum. Berlin.

Waluś, A.: Zespól narzȩdzi kowalskich i rolniczych z okresu wplywów rzymskich z


miejscowości Tluste, gm. Grodzisk Mazowiecki, woj. Warszawa, stan. 1. A set of smith’s
and farming tools of the Roman period from site 1 at Tluste, commune of Grodzisk Ma-
zowiecki. Sprawozdania Archeol. 31, 118 - 128.

Waniczek, K. 1987: Metallurgische Aussagen zur Schmiedearbeit in der Pfalz Tilleda.


Zeitschrift f. Archäol. (Berlin) 21, 91 - 111.

Wankel, J. 1882: Bilder aus der Mährischen Schweiz und ihrer Vergangenheit. Wien.

Wedel, E. von 1959: Geschichtliche Entwicklung des Umformens in Gesenken. Stahl


u. Eisen (Düsseldorf) 79, 1419 - 1427.

Weiershausen, P. 1942: Taleae ferreae. Mannus (Leipzig) 34, 84 - 92.

Weiland, H., Bunge, H.-J. 1988 [1991]: Metallkundliche Untersuchungen an Nägeln


der Römerschiffe von Mainz. In: Archäometallurgie von Kupfer und Eisen in Westeuropa,
574 - 585.

Weinand, S. Piepers, W., Wegas, M. 1967: Jahresbericht des staatlichen Ve-


trauensmannes für kulturgechichtliche Bodenaltertümer vom 1. Jan. bis 31. Dez. 1965.
Bonner Jahrbücher 167, see 435 - 438.

Weinrich-Kemkes, S. 1993: Zwei Metalldepots aus dem römischen Vicus von Walldürn,
Neckar-Odenwaldkreis. Fundberichte aus Baden-Württenberg 18, 253 - 323.

Weisgerber, G., Roden, Chr. 1985: Römische Schmiedeszenen und ihre Gebläse.
Der Anschnitt 37/1, 2 - 21.

Welkov, J. 1935: Eine Gotenfestung bei Sadovetz. Germania 19, 152 - 158.

Wensky, M. 1997: Frauen im Handwerk. In: Europäische Technik im Mittelalter,


509 - 518.
316 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Werner, G. 1981: Goldsmiths’, silversmiths’ and carpenters’ tools. In: Excavations
at Helgö vol. VII. Stockholm, 39 - 62.

Werner, J. 1969: Der Lorenzberg bei Epfach. Die spätrömischen und mittelaterlichen
Anlagen. München.

Westermann’a Atlas zur Weltgeschichte. Berlin etc. 1963.

Westphal, H. 1984: Besondere Schweisstechnik an zwei Saxklingen des 7. Jahrhun-


derts von Lembach, Stadt Dorsten. Ausgrabungen u. Funde in Westfalen-Lippe 2/1984,
57 - 68.

Westphal, H. 1991: Untersuchungen an Saxklingen des sächsischen Stammesgebietes.


Schmiedetechnik, Typologie, Dekoration. Studie z. Sachsenforschung 7, 271 - 365.

Westphalen, P. 1989: Die Schmiedeschlacken von Haithabu. Ein Beitrag zur Geschichte
des Schmiedehandwerks in Nordeuropa. Berichte über die Ausgrabungen in Haithabu vol.
26. Neumünster.

Wielowiejski, J. 1975: Kultura materialna starożytnej Grecji I. Warszawa.

Wilhelmi, K. 1977: Ein ”neuer”, zweiter Barrenfund der vorrömischen Eisenzeit


nördlich des Rheins. Germania 55, 184 - 190.

Willen, D., Soedel, W., Foley, V. 1976: The Story of Weyland the Smith. HM
10/2, 84 - 86.

Williams, A. R. 1978: A note on the manufacture of armour in 15th century Italy.


Metropolitan Museum Journal (New York) 13, 131sq.

Williams, A. R. 1987: The manufacture of armour in Germany; the metallographic


evidence from specimens in German museums. Waffen-und Kostümkunde (München) 46,
90sq.

Williams, A. R. 1991: Italian armour and Cosimo dei Medici. Journal of the Arms
and Armour Society 13/5, 243 - 315.

Williams, A. R. 1993: Slag inclusions in armour plate (1400 - 1640). In: Bloomery
Ironmaking during 2000 Years III, 115 - 121.

Williams, A. 1994: The Blast Furnace and the Mass Production of Armour Plate.
In: History of Technology 16, 98 - 138.

Williams, A. 2003: Biringuccio and the metallurgy of Italian armour. In: Ar-
chaeometallurgy in Europe (Milan), 143 - 151.
BIBLIOGRAPHY 317
Williams, A. 2004: Metallography of some 14th century European helmets. In: Met-
allography 2004. Košice, 645 - 651.

Willms, Chr. 1995: Hoch- und spätmittelalterliche Eisengewinnung an der oberen


Dill/Dietzhölze. In: Eisenland. Zu den Wurzeln der nassauischen Eisenindustrie (B.
Pipsker ed.). Wiesbaden, 57 - 81.

Wilson, D. M. 1960. The Anglosaxons. London.

Wilthew, P. 1987: Metallographic examination of medieval knives and shears. In:


J. Cowgal, M. de Neergaard, W. Griffith: Knives and Scabbards, Medieval finds from
Excavations in London 1. London, 62 - 74.

Wizcaino, A., Budd, P., McDonnell, G. 1998: An Experimental Investigation of


the Behaviour of Phosphorus in Bloomery Iron. Bulletin of the Metals Museum (Sendai)
29, 13 - 19.

Wolters, J. 1997: Drahtherstellung im Mittelalter. In: Europäische Technik im Mit-


telalter. Berlin, 205 - 216.

Wright, G. E. 1939: Iron: The Date of its Introduction in Palestine. AJA 43, 458 -
463.

Wulff, H. E. 1966: The Traditional Crafts of Persia, Their Development, Technology,


and Influence on Eastern and Western Civilizations. The M.I.T. Press; Cambridge, Mass.,
London.

Ypey, J. 1963: Vroeg-middeleeuwse wapens uit Nederlandse verzamlingen. Berichten


van de rijksdienst voor het oudheidkundig bodemonderzoek 12-13 1962/1963, 153 - 176.

Zaitz, E. 1980: Żelazne grzywny siekieropodobne z ul. Kanoniczej 13 w Krakowie.


Wiadomości Numizmatyczne 24/1, 24 - 28.

Zaitz, E. 1988: Frühmittelalterliche axtförmige Eisenbarren aus Kleinpolen. Slovenská


archeol. 36/2, 261 - 276.

Zaitz, E. 1990: Wczesnośredniowieczne grzywny siekieropodobne z Malopolski. Früh-


mittelalterliche axtförmige Eisenbarren aus Kleinpolen. Materily Archeol. (Kraków) 25,
167 - 177.

Zav’yalov, V. I. 2002: Zhelezhnyy inventar’ Polomskoy kultury (po materialam


varninskogo arkheologicheskogo kompleksa). Iron artefacts of Polomskaya culture (ac-
cording to the materials of Varni archaeological assemblage). Kratkiye soobch. Inst.
arkh. Ross. akad. nauk (Moskva) 213, 101 - 108.

Zavyalov, V. I. 2003: The knives of the Polom culture. In: Archaeometallurgy in


Europe (Milan), 165 - 171.
318 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Zav’yalov, V. I., Rozanova, L. S., Terekhova, N.N. 2001: Arkheometallografiya


kak istoricheskiy istochnik (itogi izucheniya kuznechnogo remesla v Rossii za 1995 - 2000
gody). Archeometallography as a historical source. The results of blacksmith’s produc-
tion in Russia 1995 - 2000.

Zimmer, G. 1982. Römische Berufsdarstellungen. Archäol. Forschungen vol. 12.


Berlin.

Zimmermann, Chr. 1998: Zur Entwicklung der Eisenmetallurgie in Skandinavien


und Schleswig-Holstein. Präh. Zeitschrift 73/1, 69 - 79.

Zimmermann, U. 1993: Untersuchung zum frühen Bergbau im Südschwarzwald. In:


Montanarchäologie in Europa, 201 - 229.

Zimny, J. 1965: Metaloznawcze badania wyrobów żelaznych z Czȩstochowy-Rakowa.


Rocznik Muzeum w Czȩstochowej I, 329 - 384.

Zimny, J. 1966: Halsztackie wyroby z żelaza meteorytowego z Czȩstochowy-Rakowa


(700 - 550 w. p. n. e). Z Otchlani Wieków 32/1, 29 - 43.

Ziomecki, J. 1975: Les représentations d’artisan sur les vases antiques. Wroclaw,
Warszawa, Kraków.

Zöller, S. 1997: Literatur und Technik im Mittelalter. In: Europäische Technik im


Mittelalter, 525 - 535.

Zwicker, U. 1967: Untersuchungen an einem Eisenbarren aus dem Depotfund von


Renningen (Kr. Leonberg). Fundberichte aus Schwaben N.F. (Stuttgart) 18, 282 - 283.

Zwicker, U. 1996: Ein Werkzeugstahl des Saturninus. Bayerische Vorgeschichtsblät-


ter 61, 245 - 246.

ADDENDA

Agorregiko burdinola eta errotak (Aia, Guipuzcoa). La ferreria y les molinos de Agor-
regi I - II. (M. Urteaga ed.). Guipuzcoa 2002.

Armbruster, B. R. 2005: Schmiede in der Ur- und Frühgeschichte. In: Ferrum 77,
34 - 49.

Baumeister, M. 2004: Metallrecycling in der Frühgeschiche. Untersuchungen zur


technischen, wirtschaftlichen und gesellschaftlichen Rolle sekundärer Metallverwertung
im 1. Jahrtausend n. Chr. Würzburger Arbeiten zur Prähistorischen Archäologie Band
3 (W. Schier ed.), Rahden/Westf.
BIBLIOGRAPHY 319
Biborski, M, Kaczanowski, P., Kȩdzierski, Z., Stȩmpiński, J. 2002: Bada-
nia nad technologia̧ mieczy z mlodszego okresu przedrzymskiego z obszaru kultury prze-
worskiej. In: Varia Barbarica. Monumenta Archaeologica Barbarica, Series Gemina,
Tomus I. Warszawa, Lublin, 81 - 104.

Biborski, M., Kaczanowski, P. 2003: Uwagi o technologii mieczy jako kryterium


identifikacji importów broni rzymskiej. In: Antik i Barbarzyńcy. Ksiȩga dedykowana
Profesoru Jerzemu Kolendo w siedemdziea̧ta̧ rocznicȩ urodzenin (A. Bursche, R. Ciolek
eds). Warszawa, 109 - 120.

Buchwald, V. F. 2005: Iron and steel in ancient times. Det Kongelige Danske Vi-
denskabernes Selskab, Historisk-filosofiske Skrifter vol. 29, Copenhagen.

Dungworth, D., Wilkes, R. 2005: Spherical Hammerscale and Experimental Black-


smithing. HMS NEWS 59, 3 - 4.

Espelund, A. [2005]: Jernet i Vest-Telemark det tussane råde grunnen. Iron in West-
ern Telemark where the goblins ruled. Trondheim

Ferrum 77 2005: Schmieden in Geschichte und Gegenwart. Ferrum, Nachrichten aus


der Eisenbibliothek, Stiftung der Georg Fischer Stiftung AG, Schlatt TG.

Jaritz, G. 2005: Der Schmied im Bild des Spätmittelalters und der frühen Neuzeit.
In: Ferrum 77,78 - 87.

Jockenhövel, A., Willms, Chr. 2005: Das Dietzhölzetal-Projekt. Archäometallur-


gische Untersuchungen zur Geschichte und Struktur der mittelalterlichen Eisengewinnung
im Lahn-Dill-Gebiet (Hessen). Mit Beiträgen von M. Bušs, I Keesmann, A. Kromz, D.
Lammers, R. Pott und M. Speier. Verlag Marie Leidorf, Rahden/Westf.

Kinder, J. 2005: Damaszierte Schwertklingen - Wie lange und bei welcher Temper-
atur wurden geschmiedet?. In: Ferrum 77, 99 - 112.

Knöppel, K. D. 2005: Freiformschmieden. In: Ferrum 77, 4 - 15.

Orengo, L. 2003: Forges et forgerons dans les habitats latèniens de la Grande Limagne
d’Auvergne. Fabrication et consommation des produits manufacturés en fer en Gaule à
l’Âge du Fer. Monographies instrumentum 26 (Èd. Monique Mergoil). Montagnac.

Paulinyi, A. 2005, Der Bergschmied. In: Ferrum 77, 88 - 98.

Senn-Bischofsberger 2005: Das Schmiedehandwerk im nordalpinen Raum von der


Eisenzeit bis ins frühe Mittelalter. Edition Internationale Archäologie, Naturwissenschaft
und Technologie, Band 5 (C. Dobiat, L. Laidorf eds). Rahden/Westf.

Simon-Muscheid, K. 2005: Was nutzt die Schusterin dem Schmied? Schmiedin-


nen als Abbild der verkehhrten Welt und ”reale” Frauenarbeit im mittelalterlichen und
320 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
frühneuzeitlichen Schmiedehandwerk. In: Ferrum 77m 50 - 55.

Wanhill, R. 2006: Ivestigation of a broken pile-shoe from a Roman bridge. HMS


News 61, 1.
BIBLIOGRAPHY 321

SELECTED ABBREVIATIONS

CPSA: Comité pour la Sidérurgie Ancienne de l’Union Internationale de Sciences Préhi-


storiques et Protohistoriques affilié à l’UNESCO.

DARA: Documents d’Archéologie en Rhône-Alpes, Lyon.

HM: Historical Metallurgy. The Journal of the Historical Metallurgy Society, London.

JESHO: Journal of Economic and Social History of the Orient, Leiden.

JISI: Journal of the Iron and Steel Institute, London.

KS: Kratkiye soobcheniya o dokladakh i polevykh issledovaniyakh Instituta Arkheologii,


Moskva.

MASCA: Museum Applied Science Center for Archaeology, Philadelphia.

MIA: Materialy i issledovaniya po Arkhaeologii SSSR, Moskva.

PACT: Journal of the European Study Group on Physical, Chemical, Biological and
Mathematical Techniques Applied to Archaeology. Council of Europe, Strasbourg.

SCIV: Studii şi Cercetări de Istorie Veche, Bucuresti.

Sov. arkh: Sovetskaya arkheologiya, Moskva.

ZAK: Zeitschrif für schweizerische Archäologie und Kunstgeschichte, Basel.


INDEXES

Index of places

Aachen, Germany, 40 Augsburg, Germany, 72


Aardal, Sweden, 47 Augst, Switzerland, 40, 100, 103
Abidini, Belarus, 52 Aulnut, France, 33
Abrahám, Slovakia, 199, Pl. VII Autun, France, 117, 120, 127, 133, 140,
Africa, 16, 20, 24, 50, 71, 77, 131 141, 148, 230, 244
Agri decumates, Germany, 224 Avenches, Switzerland, 117
Alaca, Turkey, 8, 226 Avestra, Sweden, 216
Alaska, America, 223 Avignon, France, 150, 238
Alésia, Alise-Ste-Reine, France, 118, 120, Awre, England, 112
140, 147-148, 305, 230, 123, 128, 146,
244 Baageroostwelde, Holland, 8
Algustrum, Sweden, 45 Bagaud, Sardinia, 43
Alsted, England, 176 Ballynahynch, Ireland, 206, Pl. XXI
Altenerdig, Germany, 216 Bargen, Switzerland, 238
Alvesta, Sweden, 84 Bargun, Germany, 77, 79
Amathus, Cyprus, 67, 227 Baudecet, Belgium, 125
Ampurias, Spain, 108, 125, 138 Bavai, Belgium, 147
Anatolia, Turkey, 6, 8, 10, 12, 17, 23, 226- Bechtheim, Germany, 28
227, 230, 234 Beckford, England, 37
Ancernant, France, 129, 132, 148-149 Belaya Vezha-Sarkel, Russia, 125, 131,
Andernach, Germany, 39 133-134, 162, 170
Appleford, Englsnd, 33 Belskoye fort, Ukraine, 153, 235
Aquileia, Italy, 95, 97, 123, 128, 146 Ben Afeli, Spain, 41
Arbrå, Sweden, 43 Berching-Pollanten, Germany 84
Arbutsro, Sweden, 157 Berlin, Germany, 84, 86, 97
Argos, Greece, 38 Bern-Tiefenau, Switzerland, 33, 72, 74
Armsheim, Germany, 26 Beroun, Bohemia, 157
Asia, 131, 229, 245-246, 248-250, 256; Asia Bešeňov, Slovakia, 216, Pl. XXXII-
Minor, 229, 234, 245-246, 248-249; see XXXIII
also Anatolia Bezdědovice, Bohemia, 74, 187
Aska, Sweden 80 Béziers, France, 134, 137-138, 151
Askra, Greece, 248 Biberist-Spitalhof, Switzerland, 121, 129,
Assum, Germany, 172 149-150, 159, 230
Assyria see Neo-Assyria Bibracte-Mont Beuvray, France, 90, 125,
Athens, Greece, 149; Agora 185; 132, 136, 155, 232, 244
Parthenon, Erechtheion, 206, 228, Pl. Bichkin Buluk, Russia, 5
V Biel, Switzerland, 189
Aubstadt, German, 26, 27 Birka, Sweden, 102
Auerberg, Germany, 139 Biskopskulle, Sweden, 43
INDEX OF PLACES 323
Biskupiń, Poland, 26, 27, 51 Carlisle, England, 143
Bixby, USA, 112 Carnuntum, Austria, 234
Blagaj-Maslovare, Croatia, 144, 145, 230 Carrier d’en Pujol, Spain, 143
Blessey-Salmaise, France, 118, 12, 148 Carthage, Alger, 24, 26, 138, 253
Bnin, Poland, 67 Čáslav-Hrádek, Bohemia, 57, 125, 200,
Bobrichtche fort, Russsia, 91 212, Pl. X, Pl. XXX
Bocholt, Germany, 166 Castellar de Pontós, Spain see Mas Castel-
Bochum, Germany, 116 lar de Pontós
Boghazköy-Hattussas, Turkey, 227, 248, Castellruf-Sainte-Marie-des-Materolles,
249 Spain, 138
Boddin, Germany, 71 Catterick Bridge, England, 65
Boldyrevo, Russia, 5, 234, 241 Caucasus, 8, 12, 234, 235, 246, 249, 254-
Bonifacio, Corsica, 41 256
Bordeaux, France, 125, 147 Čebovce, Slovakia, 49
Borodinskoye fort, Russia, 105 Čelina, Bohemia, 181
Boston, USA, 76, 84 Celle-sur-Loire, France, 73, 126, 132, 153
Bourges, France, 142 Cencelle, Italy, 166
Bourton-on-the-Water, England, 35, 37 Châbles, Switzerland, 146
Boux, France, 118, 148 Chailly en Brie, France, 143
Bram, France, 109, 125, 130, 142 Châlons-sur-Sâone, France, 189
Brampton, England, 188 Charnali, Georgia, 10
Brandes, France, 133, 179, 180 Chartres, France, 129, 143
Brandsundsnäset, Sweden, 79, 83 Châtelet, France, 102
Břasy, Bohemia, 61 Chicago, USA, 23, 25
Bredon Hill, England, 77 China, 6
Březno, Bohemia, 46, 47, 210, Pl. XXV Chios, Greece, 39
Brno-Lı́šeň, Moravia, 47, 48 Choga, Georgia, 10
Bryn, Norway, 79 Chojno near Rawicz, Poland, 204
Bryn y Castell, Wales, 116 Churburg, Austria, 223
Bucholt, Germany, 166 Citeaux, France, 246
Budeč hillfort, Bohemia, 120, 211 Clairvaux, France, 246
Bukhak, Ukraine, 222, Pl. XXXVII Cleebronn, Germany, 215, 232
Bulberry, England, 74 Clermond Ferrand, France, 155
By, Norway, 102 Coffinswall, England, 37
Byblos, Syria, 253 Colchis, Georgia, 4, 12, 246
Býčı́ skála cave, Moravia, 71, 74, 77, 81, Coppenhagen, Denmark, 118
82, 84, 84, 85, 152, 186, 245 Corbridge, England, 59, 65, 128, 143
Bygland, Norway, 74, 84, 89, 91, 102, 103 Corsica, 41
Bylany, Bohemia, 186, 201, 245 Coulmier-le-Sec, France, 30
Bytom Odrzański, Poland, 91 Cracow, Poland, 48, 49, 50, 221, 253
Byzance, Byzantion, Turkey, 39, 144, 161, Crawcwelt, Wales, 116
166, 237, 247 Cremona, Italy, 142
Crete, Greece, 10, 39, 251
Cáceres el Viejo, Spain, 40, 72, 77, 78, 87 Csatár, Hungary, 169
Camerton, England, 110 Cucuiş, Romaania, 86
Cape York, Greenland, 5 Cuvio, Italy, 215, 232
Carichin Grad, Serbia, 91 Cyprus, 18, 66, 203, 227-228, 246, Pl. XVI
324 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Dabergotz, Germany, 34 Florence, Italy, 224
Damascus, Syria, 12, 214, 246 Fontenay, France, 178
Danebury, England, 37 Forest of Dean, Wales, 37, 238
Darcey, France, 148 Fortetsa, Crete, 39
Datchet, England, 35 Fribourg, Switzerland, 117
Delos, Greece, 37 Frödinge, Sweden, 44
Delphi, Greece, 26, 27, 39 Furingstad, Sweden, 94
Dessau, Germany, 73 Furuby, Sweden, 45
Dessau-Mosigkau, Germany, 220, 236 Fyrkat, Denmark, 132, 161
Develier-Courtételle, Switzerland, 172
Dietikon, Switzerland, 117 Gailenkirchen, Germany, 184
Disco, Greenland, 20 Gajary-Pustatina Vrablicova, Slovakia,
Djikhandjuri, Georgia, 10 188
Dobřichov-Pičhora, Bohemia, 57, 58, 61 Gamów, Poland, 184
Dobřichov-Třebická, Bohemia, 62 Gannor, Gotland, 73
Dombóvár, Hungary, 61, 79, 102, 231 Gánovce, Slovakia, 8, 226
Domergue à Sauvian, France, 150 Gastein, Bad Gastein, Austria, 130, 133,
Dorak, Turkey, 8 180, 182
Dorset, England, 218, Pl. XXXV Gat’ fort, Russia, 164
Dragoslaveni, Romania, 75, 78, 82 Gáta, Slovakia, 205
Drinopol, Turkey, 142 Gaul, Gallia, 142, 147, 155, 230, 231, 244,
Dunapentele-Dunaújváros, Hungary, 26, 245, 247, 251
27 Gdańsk, Poland, 63, 221
Dürrnberg bei Hallein, Austria, 83 Gembloux-Baudecet, Belgium, 125, 147
Geoy Tepe, Iran, 8
Egypt, 5, 6, 252 Gerar-Tell Jemme, Palestine, 124
Eindhoven, Holland, 113 Germania, 59, 147, 149, 231, 247, 254
Einöd-Noreia, Austria, 128, 146, 230 Gettenau, Germany, 34
Einzing, Germany, 93, 95 Gissey, France, 129, 138
Eketorp, Sweden, 45 Glastonbury, England, 35
Ekimaucy fort, Russia, 164 Glazov, Russia, 67, 205, Pl. XIX
El Gerzeh, Egypt, 6 Glenariff, Ireland, 68
El Vilarenc, Spain, 103, 128, 138, 141, 151 Golling-Nikolausburg, Austria, 74, 83, 87,
Ely, England, 37 94, 105
Els Vilars-Arbeca, Spain, 124 Gorgippia, Ukraine, 229, 235
Elba, Italy, 134, 230 Gortys, Crete, 37, 247
Epolding-Mühltal, Bavaria, 113, 121 Gorzegno, Italy, 95, 97
Eschenburg-Wissenbach, Germany, 115, Gotland, Sweden, 247; see also Mästermyr
129, 174 Gournay sur Aronde, France, 190, 254
Eski Kahta, Turkey, 114 Gradiştea Muncelului see Sarmizegetusa
Eurasia see Chapter I: 3-9 Grässgård, Sweden, 45
Great Moravia, 118, 161, 162, 188,209,
210, 247, 237, 251, 290
Fagerhult, Sweden, 44 Greece, Greek, 5, 10, 14, 18, 37-39, 38, 50,
Fayal, Azores, 240 72, 75, 76,77, 81, 126, 131, 133, 135,
Feurs, France, 132, 155 138, 185, 190, 191, 206-229, 231, 246-
Flavigny, France, 118, 148 248, 253, 255
INDEX OF PLACES 325
Greenland, 173 Hradištko-Sekanka, Bohemia, see Sekanka
Grenoble, France, 245 Hrádok near Trenčı́n, Slovakia, 47, 48
Gretton, England, 37 Hrazany oppidum, Bohemia, 132, 152,
Grönbjerg, Seden, 157 156, 232
Gross Raden, Germany, 92 Hromówka, Poland, 214, 216
Gudbrandsdalen, Norway, 45 Hrvatska Dubica, Croatia, 41, 42
Hrynchuk fort, Ukraine, 164
Habrůvka-Padouchov, Moravia, 132 Huckhoe, England, 127, 157
Hadrianopolis, Drinopol, Turkey, 142
Haglebuvand, Norway, 74 Iceland, 93, 94, 173
Haithabu, Jutland, 50, 57, 113, 121, 132, Idalion, Cyprus, 66, 227
160, 205-206, 212, 219-220, 248 Pl. XX; Illerup, Denmark, 190, 216, 254
XXIX; XXXVI Inchtuthil, Scotland, 128, 143, 188, 231
Hällby, Sweden, 152 India, 250
Halle, Germany, 220 Innsbruck, Austria, 224
Hallstatt, Austria, 40, 84, 99, 186, 248 Ischia, Pithekoussai, Italy, 126, 135, 138,
Hallunda, Sweden, 152 229
Ham Hill, England, 35 Ivanovice, Moravia, 56, 208, 210-211, Pl.
Hama, Syria, 10 XV; XIX
Hamburg, Germany, 161, 254
Hamwic, England, 51, 236 Jarlshof, Shetlands, 173
Haus Rhade, Germany, 170, 174 Jarohněvice, Bohemia, 56
Hayvoron, Ukraine, 252 Jaroměř, Bohemia, 217
Heek-Nienborg, Jutland, 158 Jemelle, Belgium, 150
Heidetränk oppidum, Germany, 33, 158 Jenišův Újezd, Bohemia, 204, Pl. XVII
Heiligenkreuz, Austria, 114 Jindřichův Hradec, Bohemia, 175
Heiligenstadt, Germany, 34 Joa, Norway, 99
Helgö, Sweden, 43, 44, 45, 133, 205, 248, Joldelund, Germany, 116, 134, 158-159
256, Pl. XVIII Jordan, 198, 227
Helmershausen, Germany, 256 Jouars Pontchartrain, France, 94
Helpfau, Germany, 184
Heuneburg, Germany, 26, 27 Kaisheim, Germany, 28, 59
Hidészég, Hungary, 169 Kalisz, Poland, 73
Hillsborough, Ireland, 202, Pl. XI Kamionka Nadbuźna, Poland, 174
Hjortspring, Jutland, 189 Kapalica, Poland, 157
Hod Hill, England, 35 Kappel, Germany, 74, 79, 83, 87, 189
Hodde, Jutland, 154 Kapušany, Slovakia, 177
Hökbäck, Sweden, 43 Karabash, Serbia, 90, 92
Holiare, Slovakia, 8 207 Karmir Blur, Iraq, 24
Holme Chase, England, 35 Kelheim, Germany, 250
Holubice, Moravia, 203, Pl. XIV Kerch, Ukraine, 229
Holy Cross Mountains, Poland, 113, 157, Kervennenec see Pont-Croix
234, 249, 253 Khorsabad, Iraq, 12, 23, 25, 227, 250
Homburg v. d. Höhe, Germany, 30 Kilberg, Ireland, 196-197, PL. IV
Horath, Germany, 127, 150, 157 Kiyv, Ukraine, 164, 235, 255
Hostýn oppidum, Moravia, 201, Pl. XII Kjøldbymagle, Sweden, 184
Houghton Down, England, 36 Kjula, Sweden, 207-208, Pl, XVIII
326 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Klöckeberga, Sweden, 45 Lemnos, Greece, 247-248
Kletsk, Belarus, 205, Pl. XVIII Les Sorres, Spain, 43
Klučov, Bohemia, 113, 197, P. III Lhotice (Hradiště u Lhotic) oppidum, Bo-
Knyazha Gora fort, Ukraine, 87 hemia, 152, 232
Koblemz, Germany, 250 Liberec, Bohemia, 118
Kobyly, Bohemia, 186, 250 Libice, Bohemia, 217
Količı́n, Moravia, 14, 56 Liestol-Munzach, Switzerland, 169
Kolı́n, Bohemia, 61, 187, 216-217 Linga, Sweden, 151
Köln, Germany, 147 Liptovská Mara oppidum, Slovakia, 100
Kolyu, Bulgaria, 78, 89 Littleton, England, 35
Korytnica, Poland, 73, 88 Ljungby, Sweden, 44
Korchak, Ukraine, 253 Llanstephan, England, 35
Kosel, Germany, 133, 169, 170 Llyn Cerrig Bach, England, 35, 59, 63,
Košice, Slovakia, 117, 195 189, 211, 251, 254
Kostolná, Slovakia, 199-200, Pl. IX London, England, 76, 84, 135, 165, 202
Kraków see Cracow Lovosice, Bohemia, 57, 61, 67, 201, Pl. XI
Kreimbach, Germany, 75, 77, 78, 93, 94, Lozna, Romania, 74, 83
98, 105, 106, 231 Lübeck, Germany, 161, 254
Kriftel, Germany, 127, 147 Lubovňa castle, Slovakia, 177
Krivodol, Bulgaria, 82 Lucca, Italy, 142
Krokerud, Norway, 47 Lutomiersk, Poland, 63, 212, Pl. XXVIII
Kundl Lus, Tirol, 116, 154-155 Lyon, France, 132
Kung Chien, China, 6
Kúty, Slovakia, 61 Mâcon, France, 142
Kyjatice, Slovakia, 198, 280 Magdalensberg, Austria, 23, 24, 40, 42-43,
72, 77, 78, 92, 93, 95, 99, 100, 103, 127,
154, 192, 251-252
La Tène, Switzerland, 10, 56, 57, 59, 89, Magdeburg, Germany, 220
250, 254 Maidenhead, England, 35
Lahovice, Bohemia, 220 Mainz, Germany, 90, 104, 105, 115
Landshut, Austria, 223-224 Malaya Kopanya, Ukraine, 117
L’Ans-aux-Meadows, New Foundland, Malvern, England, 35
173 Manching, Germany, 23, 24, 28, 33, 68,
Lapithos, Cyprus, 66 71, 83, 84, 87, 91, 98, 103, 115-116,
Lapphytan, Sweden, 238 122, 126, 132, 155, 192, 209, 232, 251,-
Latium, Latin, Italy, 18, 124-125, 130, 132 252, Pl. XXIV
Lattes, France, 124-125, 130, 137, 138, 148 Marculani, Romania, 74-75
Laurion, Greece, 206, 229 Marseille, Massalia, France, 126, 133, 137-
Lausanne, Switzerland, 116 138
Lebedka, Ukraine, 128, 169, 170 Marsens, Switzerland, 117, 116, 130
Legva, Georgia, 10 Martigues, France, 126, 134, 136, 137-138
Leipzig-Thekla, Germany, 73 Mas Castellar de Pontós, Spain, 109, 125,
Leipzig-Wahren, Germany, 26, 27 134, 136, 138
Lejre, Denmark, 73, 103 Maslovare see Blagaj-Maslovare
Lelekovice castle, Moravia, 117, 133, 175, Massalia see Marseille
176 Mästermyr, Gotland, 50, 51, 75, 97, 101,
Lemeshevichi, Belarus, 212 104, 251
INDEX OF PLACES 327
Maszkowice, Poland, 26, 27 Newstead, Scotland, 74, 231
Matteille, Sardinia, 43 Niederzier-Hambach, Germany, 43
Mayatsk fort, Ukraine, 236, 250, 254 Nı̂mes, France, 142, 147
Mayran, France, 150 Nimrud, Iraq, 23-24, 25, 26
Menka, Belarus, 208, Pl. XXI Nitra-L’upka, Slovakia, 211, Pl.XXVIII
Menzlin, Germany, 122, 161 Nord Roldnes, Norway, 102
Meon Hill, England, 35 Nordre Besseberg, Norway, 87
Mezöbánd (Marostorda), Romania, 73, 98 Noreia see Einöd-Noreia
Mikulčice, Moravia, 47, 162, 163, 188, Noricum, Norician, 154, 231, 251, 252
204, 210, Pl.XXVI Nörten-Hardenberg, 115
Milan, Italy, 224, 238 Nosaby, Sweden, 50, 51
Miletus, Turkey, 248 Nottingham, England, 165
Minepit Wood, England, 127, 154 Novgorod Velikiy, Russia, 61, 63, 72, 79,
Mont Auxois see Alesia 92, 95, 98, 102-103, 195, 220
Mont Beuvray-Bibracte, France, 125 Nowa Boćwinka, Poland, 88
Mont Lassois, France, 28 Nowa Huta-Mogila, Poland, 211
Montans, France, 33, 36 Nuits-Saint-Georges, France, 145
Moosberg, Germany, 78, 83, 92, 94 Nuremberg, Germany, 164, 224, 238
Mörigen, Switzerland, 184 Nydam, Jutland, 190
Moravský Ján, Slovakia, 54
Morro di Mezquitilla, Spain, 115
Mšec, Bohemia, 109, 136, 154 Oberwinterthur, Switzerland, 42, 70, 117,
Mt. Adir, Palestine, 66 139
Mshvidobauri, Georgia, 10 Oblin, Poland, 216
Munich, Germany, Pl. XXXIX Ochtrup, Germany, 33, 35, 106
Münsingen, Switzerland, 211, 250 Offerdal, Norway, 43
Münster, Germany, 33, 34 Öland, Sweden, 44, 45
Mutějovice, Bohemia, 121, 170, 171 221, Olbia, Ukraine, 39, 235
225 Olynthus, Greece, 77
Mycenae, Greece, 10 Opole, Poland, 132, 221
Mysli, Belarus, 52 Orton, England, 35, 37
Mziani, Georgia, 10 Orvieto, Italy, 135
Ošaniči, Herzegovina, 74
Nailly, France, 111, 112, 117, 130 Osnabrück, Germany, 217
Namur, Belgium, 33, 238 Ostprignitz, Germany, 17
Nancy, France, 118 Ovifak, Greenland, 20
Naukratis, Egypt, 26, 27
Navasiolki, Belarus, Pl. XXV Palavas, France, 41
Neftenbach, Switzerland, 117 Paleopaphos-Skales, Cyprus, 66
Nejdek, Moravia, 47, 48, 201 Pampailly, France, 110, 180, 181-182, 208,
Nemeskér, Hungaria, 172 Pl. XXIII
Neo-Assyria, realm of, 12, 23-24, 25, 26, Pantikappaion, Ukraine, 235
227. 246, 252, 256 Park Farm, England, 36
Netolice, Bohemia, 258 Pasargadae, Iran, 197, 244
Neunkirchen-Zeppenfeld, Germany, 154 Pastyrskoye gorodichtche for, Ukraine, 94,
Newbury, England, Pl. XXVIII 212, 252
Niedererlbach, Germany, 115 Pella, Jordan, 198, 226-227
328 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Pendjikent, Uzbekistan, 164 Radzovce, Slovakia. 198, 227
Pen’kivka, Ukraine, 235, 252 Ramières, France, 109, 133, 150
Perachora, Greece, 37, 39 Rayki fort, Russia, 91
Persia, 231, 244, 246, 257 Reask, Ireland, 178
Persepolis, Iran, 197-198 Redikar, Russia, 73
Petöháza, Hungary, 127, 151, 230 Regensburg, Germany, 115, 234
Petrovo, Ukraine, 117 Renningen, Germany, 28, 30
Phanagoria, Ukraine, 235 Reykjaqvı́k, Iceland, 132, 173
Phokis, Greece, 39 Rheinau, Switzerland, 127, 132, 155, 232
Piekary, Poland, 51 Rheinzabern, Germany, 133
Piotrawin, Poland, 48, 49 Rheneia island, Greece, 37
Piraeus, Greece, 229 Ribe, Denmark, 132, 166, 173
Pithkoussai see Ischia Riom, Switzerland, 128, 150
Plavecké Podhradie, Slovakia, 30 Risegjerdet, Norway, 91
Pleets, Germany, 73 Roanne, France, 132
Plemirio, Sicily, 43 Rocca di San Silvestro, Italy, 166
Pliska, Bulgaria, 166, 253 Rodalben, Germany, 28
Plotiště, Bohemia, 216 Rom, France, 143
Pobedim, Slovakia, 48 Rome, Italy, 72, 123, 132, 228, 231-232,
Podbolot’e, Russia, 67, 91 247, 255
Podhradı́, Moravia, 61 Romfoghellen, Norway, 104
Pohansko site, Moravia, 46, 97, 101, 139, Rumhult, Sweden, 44
141, 150 Runder Berg, Germany, 75, 99
Pokrzywnica, Poland, 157 Ruscino-Rousillon, France, 134, 142, 188
Pompeii, Italy, 97, 139, 230, 253 Ryazan’, Russia, 67
Ponětovice, Moravia, 57 Rza̧dz, Poland, 73
Pont-Croix - Kervennenec, Brittany, 129,
134, 141, 150, 230 Saalburg, Germany, 54, 91, 99
Populonia, Italy, 84, 88 Sadovec, Bulgaria, 72, 84, 98, 104, 105,
Pornullbecken, Finland, 160 144
Porolissum, Romania, 73 Saeding, Denmark, 173
Port-Nidau, Switzerland, 73, 36, 189 Šafárikovo see Tornal’a
Poysdorf, Austria, 73 Saffig, Germany, 53
Prague, Bohemia, 1, 135, 165, 202, 220, Saint Agnan, France, 93, 131-132
224, 236, 253, 256, 258 Saint-Croix-aux-Mines, France, 134, 182
Près Bas, France, 151 Saint-Jean-Trolimes, France, 30
Preslav, Bulgaria, 82, 166, 253 Saintes-Maries-de-la-Mer, France, 41, 42
Priene, Turkey, 229 Salmonsbury, England, 36
Přı́tluky, Moravia, 188 Saltovo (Verkhne Saltovo) fort, Ukraine,
Przeworsk, Poland, 186 236, 250, 254
Puissaye, France, 153 Salzach river, Austria, 51
Samarra, Iraq, 6
Rabestens, France, 33 Sanda, Sweden, 316
Råby, Sweden, 45 Sandnes, Greenland, 173
Racia̧że, Poland, 175 Sandy, England, 74
Radovanu, Romania, 94 Sanzeno, Italy, 72, 78, 87, 88, 89, 92, 94,
Radymno, Poland, 48 103, 106
INDEX OF PLACES 329
Saône river, France, 211 Staraya Ladoga, Russia, 72, 99, 102, 104
Sardinia, France, 30, 43 Staré Hradisko, Moravia, 57, 61, 63, 102
Šariš castle, Slovakia, 177 Staré Město, Moravia, 47, 48, 162, 163
Sarkel see Belaya Vezha Steinsburg-Kleiner Gleichberg oppidum,
Sarmizegetusa, Romania, 75, 82, 125, 156, 30, 61, 100
233, 246, 254 Stöng, Iceland, 130, 134, 170, 173
Sauggart, Germany, 28, 29 Stradonice oppidum, Bohemia, 61, 68, 97,
Sauvian-La Domergue, France, 40, 42, 159 101, 106, 201, 255, Pl. VII
Schwadenau, Switzerland, 30 Strasbourg, France, 28
Sekanka, Hradištko-Sekanka, Bohemia, Stuttgart, Germany, 222
213, 214, 221-222, Pl. XXXI; Pl. Süderschmedeby, Germany, 158
XXXVIII Sulzburg, Germany, 179
Selz, Germany, 75, 84, 91, 231 Sundsvall, Sweden, 43
Semice, Bohemia, 56 Susa, Iran, 23, 24
Sendražice, Bohemia, 67 Szaniec, Poland, 73, 95, 99
Sens, France, 93, 96 Szwajcaria, Poland, 67, 71
Sévaz-Tudingen, Switzerland, 153
Sezimovo Ústı́, Bohemia, 129, 130, 163,
167, 202 Taanach, Palestine, 66, 227
Sidon, Syria, 253 Tarnówek, Poland, 67
Siedla̧tków, Poland, 175 Taymanova, Belarus, 52
Siemianice, Poland, 79 Tchernyakhov, Ukraine, 235, 247, 255
Sigtuna, Sweden, 102 Tell Asmar, Turkey, 226
Silchester, England, 103 Tell el Amarna, Egypt, 252
Sincraieni, Romania, 6 Tell-el-Farah, Palestine, 66
Singelfingen, Germany, 220 Tell Halaf, Iraq, 226
Singen, Germany, 199, 228, Pl. VIII Tell Qiri, Palestine, 65
Sisak, Croatia, 41 Thasos, Greece, 229
Skedstad, Sweden, 44, 45 Thebes, Egypt, 39, 226
Skogar, Iceland, 93, 94 Thespiai, Greece, 39
Skredtveit, Norway, 73, 83 Thessaloniki, Greece, 248
Sládkovičovo, Slovakia, 201-202, Pl. XIII Thorsberg, Denmark, 190, 254
Slobodka, Belarus, 205, Pl. XVIII Tiddington, England, 110, 125, 143
Smiss, Gotland, 88 Tilleda, Germany, 161
Snorup, Jutland, 49, 155, 158, 159 Tišice, Bohemia, 67
Sobotka, Poland, 214 Tjele, Jutland, 75, 79, 87, 97, 99, 101,
Sombernon, France, 118, 148 102, 104
Souse, Tunisia, 41 Töging, Germany, 184, 185
Sparta, Greece, 38 Toledo, Spain, 164, 248
Sponeck, Germany, 96, 99 Tonsberg, Sweden, 168
Sremska Mitrovica, Serbia, 89, 92 Toome Bar, Ireland, 68
St. Gallen, Switzerland, 177 Tornbotten, Sweden, 44
St. Georgen, Austria, 72, 87 Toscanos, Spain, 114
St. Marcel, France, 33 Tossal el Moro, Spain, 151
St. Martin, Switzerland, 139 Toufreville, France, 36, 43
St. Petersburg, Russia, 104 Toulon-sur-Allier, France, 132
St. Ulrich-Bickenberg, Germany, 179 Transrand, Sweden, 74
330 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
Třebohostice, Bohemia, 196-197 Vimose, Denmark, 75, 190, 216, 254
Trenčı́n, Slovakia, 48 Vitebsk, Belarus, 222, Pl. XXXVII
Trier, Germany, 231 Vitteux, France, 148
Trondheim, Norway, 128, 168 Vladimir, Russia, 79
Tronsberg, Norway, 168 Vlastislav, Bohemia, 100
Trosky castle, Bohemia, 118, 132, 176-177 Vorbasse, Jutland, 159, 172
Troy, Turkey, 8 Voronesh, Ukraine, 8
Tunisia, 41 Wa̧chock, Poland, 214
Tvarožná Lhota, Slovakia, 210 Wald a. d. Alz, Germany, 184, 185
Tyniec, Poland, 178 Waltham Abbey, England, 74, 178
Tyras, Romania, 235 Wartburg castle, Germany, 33, 34, 36
Tyrus, Syria, 253 Waschenberg, Austria, 120, 133, 152
Weald of Sussex, England, 238
Ugarit, Syria, 213 Wérimont, Belgium, 33
Uherský Brod. Moravia, 56 Wesólki, Poland, 73, 79
Ulaka, Slovenia, 132, 141, 144 White Fort, Ireland, 203, Pl. XV
Unterkrumbach, Germany, 184, 185 Wichdorf, Germany, 34
Unterregenbach, Germany, 220 Wicina, Poland, 81
Ur, Iraq, 226 Wilderspool, England, 110, 125, 143
Urach, Germany, 75 Winchester, England, 220, 256
Urartu, Iran, 12, 23, 256 Witów, Poland, 26, 27
Wlostowice-Pulawy, Poland, 73, 89
Val de Gabbia, Italy, 118 Wroclaw, Poland, 221
Valbo, Sweden, 43 Wymyslowo, Poland, 103
Vardaroftsa, Macedonia, 10
Vatican, Italy, 123 Xanten, Germany, 80, 116
Velim, Bohemia, 90
Vel’ký Klı́ž, Slovakia, 47, 48 Yablonovka, Ukraine, 84, 85
Vendel, Sweden, 45, 70, 80, 88, 89, 234, York, England, 116, 121
248, 256
Venice, Italy, 244 Zadowice, Poland, 73
Ventlinge, Sweden, 44 Zajȩczkowice, Poland, 113
Vernay see Pampailly Zarubincy, Ukraine, 235, 257
Verrey-sous-Salmais, France, 148 Žatec, Bohemia, 167
Vestre Toten, Norway, 47 Závist oppidum, Bohemia, 125, 136, 156,
Věteřov, Moravia, 63 232
Vienna, 234, PL. XXXIX Zawada Lańckorońska, Poland, 48, 49
Vik, Norway, 91 Žitavská Tôň, Slovakia, 57
Villand, Norway, 101 Zug, Switzerland, 168
Villy-en-Auxois, France, 148 Zürich-Alpenquai, Switzerland, 184, 185
INDEX OF SUBJECTS 331
Index of subjects

Achaemenids, 197, 244 aššium, iron, 4


adamas, steel, 19 auger, 74, 188, 192, 204, 209
adze, 144, 146, 187-188, 209, 210 Avars, 144, 207, 244
Aeschylus, 6, 229, 244-245 austenite, 20, 51, 58, 76, Chapter XI: 194-
agricultural implements, 10, 184, 187, 210 195; 240, 242
air-inlet, air-duct, 119, 125, 126, 129, 132, awl, 235, 251
142, 149, 153, 159, 160, 164, 169, 177, axe, axehead, 14, 39, 55, 58, 60, 64, , 74-
241 75, 99, 154,, 162, 185-186, 188, 192,Pl.
alloy steel, 22 IV, XXI, XXII, XXIV
Amelungenlied, 70, 244, 256 axe-shaped bars, 45-49, 47, 49, 60, 69, 99,
amulet, amulets, 14, 34, 191, 228 162, 163, 185-186, 188, 192, 205-209,
amŭtum, iron, 4 210, 226-228, 232, 234, 237
Ananino culture, 152 ayah, syama ayah, iron, 48
AN.BAR, iron, 5, 8, 20
AN.BAR.GE6 , meteoritic iron, 5
anchors, 37, 74 bainite, 51, 69, 81, 195, 198, 208, 240, Pl.
Anglo-Saxon period, 244, Pl. XXXV I, XXIII
an.na, iron, 4 balances, 35, 40-41, 75, 192
annealing, Chapter V: 66, 240 bar, bars, 5, 7, 9, 11, 12, 13, 23, Chapter
Antes, Antai, eastern Slavs, 252 IV: 23-52, 31, 32, 33, 39, 51, 54-55, 75,
anvil, anvil position, bed, 54, 55, 58, 71, 88, 116, 120, 122, 146, 151-152, 154-
73-75, 84, Chapter VI: 93-97, 94-96, 155, 161, 167, 172-173, 179 184, 191-
98-100, 104-105, 110, 112, 123, 129, 192, 196, 198, 212, 231-232
133, 135-136, 139-140, 145, 149, 151- barzel, iron, 8
152, 154-159, 161, 163, 164, 166-169, battle-axe, 62, 187, 209-210, Pl. XXVI
170, 171, 173, 175, 176, 180, 181-182, bellows, 105, 123, 125-127, Chapter
191, 193, 230 VIII:131-133; 135, 141, 164, 171, 175,
anvil stone, stone anvil, 90, 142-143, 151- 177, 181, 256
152, 154, 156-159, 161, 169, 173 bellows and hearth protection, 126, 128,
Archaic period in Greece, 17, 37 133-135, 140, 151, 160-161, 169, Chap-
armament, 186-187, 190, 192 ter VIII: 121-133, 140, 144, 154, 160-
armour, 10, 14, 39, 64, 175, 190, Chap- 161, 169
ter XI: 222-224, 237, mail 189, 198, billet, 51, 52, 24, 59, 116, 119, 240
222-224, 232: plate 223, 238, 252, Pl. bipyramidal bars, ingots, 12, Chapter IV:
XXXIX 23-32, 25-29, 31,, 39-40, 42, 59, 71, 152,
armourer, armourers, 142, 164-165, 168, 232-233, 245
223. 229, 230-231, 236-238, 247, 254 Biringuccio, V. 180, 183, 244-245
arrow, arrowhead, 58, 61, 73, 122, 171, blank, blanks, 34, 36, 43, 44, 50, 52, 240
175-177, 188-189, 190, 228, 230, 235, blast furnace, 224
Pl. XXVIII bloom, blooms, 1, 3, 5, 14, 7, 9, 11, 12,
arrowsmith, 165, 244 23, 30, 51, 59, 93, 112, 114-120, 128,
arsenic in iron, 21, 50-51, 64, 114-115, 242 133, 145, 152-155, 157-159, 161, 164,
332 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
168, 172-174, 177, 196-198, 233, 238, charcoal ash, Chapter VII: 109-122; 123-
240-242 124, 182, 133-139, 143, 151, 153, 167,
bloomery, bloomeries, bloomery smelters, 169, 178, 242
slags 10, 14, 16, 19, 30, 109-110, 112- Charlemagne, 237
113, 116-120, 127, 153-154, 157-158, chemical analysis, composition, 28, 37,
161, 173, 197, 226, 236-238, 240-241, 113-118, 120, 200-204, 206, 209, 211
243, 253 chimney hood, 131, 162
booty see spoils chisel, chisels, 10, 14, 54, 58, 64, 74-77,
borax, 58, 110, 120 90,91, 99, 146, 152, 166, 175, 178, 181,
Bottichgebläse see trompe 186-188, 209
bridle bit, 74, 165, 185, 187-188, 191 chronology see dating
Bronze Age, 3, 6, 14, 71-77, 90, 151-152, Cimmerian, Cimmerians, 10, 17, 233-234,
185, 186, 198, 223, 226-227, 245-246, 246
251 Cistercians, 177, 256
burgus, 147
clamps, 14, 55, 154, 181, 191, 192, 197-
Bylany culture, 64, 186, 201
198, 206, 229, 238, Pl. V
Classical period in Greece, 17, 37, 77, 229
clock making, 225
Caesar, C. I, 32, 34
cold working, hammering of iron, 54, 123,
Carburization of iron, 20, 23-24, 26, 28,
224-225, 231, 238
30, 37, 46, Chapter V: 54-71, 90, 165,
Chapter XI: 200-202, Pl. IX, XI-XIII, colour of heated iron, 53-54, 59, 69
XXVI copper slag, 3
corrosion, 18, 35, 45-46, 64, 114-115
Carthusians, 177, 245
coulter, 34, 188, 192, 203
cartwright, 64, 139
cow bells, 142, 166, 192
Cassiodorus, 236
corrosion, 18, 35, 45-46, 64, 114-115, 149
cast iron, 8, 51, 118-119, 148, 153, 194,
238, 240-241 cross-bow, cross-bow bolt, 165, 168, 175,
casting mould, 73-75 177, 223, 238
cattle bell see cow bell crucibles, 67, 137, 144, 146, 150, 153, 137,
Celtic, Celts, 14, 23, 28, 29, 30, 32, 39, 57, 146, 153, 161, 164-166, 175, 177-178
59, 64, 105, 115, 125-126, 132, 135, 144, cuirasse, 223, 231
153-155, 189-190, 196, 198, 201, 203-05, cuneiform texts, records 7, 8, 9, 13, 15
209, 232, 244-246, 248, 251, 254, 257 currency bars, 34-37. 36, 44, 46,
cementation see carburization, Pl. VII, cutlers, cutlery, 14, 144, 144, 164-165, 169,
IX 190, 192, 205-206, 210-211, 229, 231,
cementite, 20, 40, 41, 45, 54, 66, 69, 194, 237-238, Pl. XXV, XXXVIII
195, 240, 246, Pl. VI, VII Cypriot, Cyprogeometric period, 3, 246
chain, chains, 55, 75, 146, 186-189, 191,
192
Chalybes, 6, 12, 228, 245, 257 Dacians, 75, 82, 125, 135, 153, 233, 235,
chalyps, steel, 19, 245 245-247, 255
charcoal, 10, 67, 84, 105, 109-110, 115-116, Dactyli, 6, 228
119-120, 123-126, 128-129, 130, 133- dagger, daggers, 4, 5, 8, 12, 58, 66, 174,
135, 140, 141, 146, 151, 153, 156, 164, 186-187, 189, 203, 226, Pl. XVI
167, 169, 173, 178, 181-182, 200, 224, damascene steel, damast, 60, 68, 211, 213,
241 Pl. XXXI
INDEX OF SUBJECTS 333
dating, 3, 6, 16, 20; archaeological, 3, 10, flux, 58, 109, 112-113, 115-118, 120, 181,
archaeomagnetic, 3, 10; radiocarbon 3, 240-241
10, 17, 151-153, 157, 159, 179 forge see smithy
decarburization, 37, 124 forge welding see welding
Demosthenes, 220, 246-247 forge spoon, 75, 106-107, 154-156, 187-
die, dies, 64, 71, 74-75, Chapter VI: 92-99, 188, 230
101, 104, 211, 230, see also swages forging operation, Chapter V: 53-70
Dionisius of Halicarnassus, 230, 247 fork, 58, 192
domeykite, 115 forming of iron, 53-65, Pl. V
drachme, 37 founders, founding, 71, 168
drawing iron, die, hole see wire drawing fuel, 109-110, 123, 134, 181, 183
drift, drifts, 90, 92

Gallo-Roman period, culture, 75, 132,


Edda, mythological poems, 256 140, 148, 233
equilibrium diagramme, 21, 20, 22, 53, Gaul, Gallic, 34, 64, 149, 155, 244, 252
240 gelso, gelžis, iron, 19
Etruscan, Etruscans, 10, 17, 84, 138, 229-
Geometric period in Greece, 10, 17, 37,
230, 247
228, 247
eutectoid steel, 20, 200, 240
Gepidic, Gepidi tribe, 73
experimental activities, experiments, 2,
Gerätgeld, 50
113-114, 116, 132
Germanic tribes, folks, 73, 156, 160-161,
190, 231, 233, 247, 251, 254-255
frabricae armorum, 142 Getian tribe, Getes, 233
fanum, 149, 247 glass, glassy, 111, 240
farrier, farriers, 165, 167, 171-172, 175, Gortyna laws, 37, 247
238 Goths, 144, 188, 235, 247, 252, 255
fayalite, fayalitic, 112, 115, 119-120, 154, graphite in iron, 20, 194, 240
240, 243 grid, 192
ferrite, ferritic, 20, 24, 37, 40, 48, 50, gromps, gra̧pie, 149, 240
53, 69-70, 81, 105, 148, 195, Chap- gun-balls, 193
ter XI: passim, 240, Pl. I-III, V, VII-
XI,XIV-XVII, XX, XXII, XXIV-XXX,
XXXIII-XXXIV, XXXVII, XXXIX Haedui tribe, 244
ferrum, iron, 18-19 haematite, 110, 112, 117
fetters, 192, 251 halberd, 192
fibulae (brooches), 55, 156 Hallstatt period, culture, 3, 12, 14, 17,
file, files, filing, 71, 73-76, 99-101, 103, 26,27, 55, 57, 58, 60, 62, 64, 66,-67,
237, 155-156, 166, 186, 192, 224, 256 71, 82, 85, 94, , 126, 132, 152-153, 204,
file maker 192 227, 233, 245, 247-248, 253, Pl. XI
finery, fining of steel, 224 hammer, 2, 48, 55, 58, 64, 71-81, 80-83,
fire-dog, 14, 39, 74, 189, 191, 192, 228 90, 96, 97, 100, 123, 139, 144, 146, 152,
fire-steel see strike-a-fire 155-156, 175, 181-182, 187-188, 193,
fitting, fittings, 10, 21, 54-55, 58, 174- 198, 200, 211, 230, 232, 245
175, 177, 182, 187-188, 174-175, 188- hammer-axe, 58
189, 192, 196 hammer-mill, 58, 178, 193, 197, 238-239
334 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
hammer scale, 58, Chapter VII: 110-112; iconographical sources concerned with
111, 115-120, 125, 130, 135-137, 140, iroworking, depictions, illustrations76,
150, 153-156, 158-160, 162, 163, 166- 81, 84, 86, 88, 95, 97, 104, 128, 131,
167, 171, 173-175, 181, 199, 202, 239- 133,
241, 243 Iliad, 228, 249
hapalki, habalginnu, iron, 4, 19 Illyrian tribes, 249
hardy, 191 Indo-Europeans, Indo-European, 14, 18-
harness, 193 19, 144, 227, 246, 249-251, 254
Hattians, 238 inclusions see slag inclusions
Hattusil, 227 ingot, 7, 9, 13, 29, 31, 33, 36, 41, 42, 52,
hearth, 105, 109, 111, 113-116, 118-119, 71, 119, 154, 184, 192, 232-233, 241
121, Chapter VIII: 123-131, 130, 133, inlays in iron, 14, 120, Chapter X: 184-
Chapter IX: 135-184, passim; 200 185; 185-186
heat treatment of steel, Chapter V; 65-70;
Inuit Eskimos, 5, 20
181, 200, 208-209, 216-217, 220, 224,
iron, passim
236, Pl. XXIX
Iron Age, 5, Chapter II: 10-16; 113, 135,
Hecathaeus of Milet, 248
151-152, 156, 250-251
Helladic period, 3, 14, 17, 248
Hellenistic period, 229, 248, 255 iron bars see bar. bars
helmet, 73, 175, 186, 189, 191, 223-224, iron currency, Chapter IV: 32-39; 162
230, 238, Pl. XXXIX iron stock, Chapter IV: 24-52, see also
bars, ingots
Helvetii tribe, 189, 232
ironworks see bloomeries
Hephaestus, 76, 81, 84, 86, 247
isern, iron, 18
hercynite, 119
iscorite, 119
Herodotus, 39, 248, 254
Italic tribes, Italics, 19, 138
Hesiod, 135, 185, 228, 248
Hittite, Hittites, 3, 5, 8, 20, 184, 227, 248-
249
hoard, hoards, 1, 26, 33-34, 45-46, 47, 48- jewellers, 238
50, 74-75, 77, 79, 82, 85, 105, 139, 144,
155-156, 162, 163, Chapter X: 184-193;
208, 231-232, 249, 251, 254 kalayasa, iron, 18
hoe, hoe makers, 162, 188, 190, 191 kamacite, 19, 241
Homer, Homeric epics, 228, 246, 249 kettle hooks, 187, 192
horloge, 225, 238 key, keys, 55, 74-75, 139, 142, 151, 132,
horse bit, gear, 55, 73, 61, 188, 190, 192- 187-188, 225
193, 201 Khazar tribes, 162, 164, 236, 350, 254
horseshoe, 146, 167, 171-172, 175, 177, kirschsteinite, 119
182 Kobyly culture, 186, 250
Hussites, 167, 224 Korchak culture, 250
hypereutectoid steel, 20, 40, 67, 100, 148, knife, knives, 6, 10, 12, 14, 51, 60, 64, 66,
198, 202, 240, 246 68, 73, 122, 146, 150, 165, 167, 174,
hypocaustum, 151 185-189, 191-192, 198, 200-201, 203,
205, 209, 211-212, 215, 220-222, 227-
228, 233-236, 246, Pl. VI, IX, XI-XIII,
Iapodi, Iapudes, Illyrian tribe, 144 XVIII-XIX, XXI, XXV, XXVII, XXX-
iarn, isarnon, isern, iron, 18 XXXI, XXXV, XXXVII-XXXVIII
INDEX OF SUBJECTS 335
kù.an, iron, 4 XXV, XXVIII, XXX, XXXI, XXXIV-
Kyjatice culture, 198, 250 XXXIX
melilithe, 119
Mendelsche Zwölfbruderstiftung, 104
La Tène period, 12, 14, 17, 26, 27, 28, 31, metallography, metallographic investiga-
36, 55. 56, 60-61, 68, 72, 74, 77, 79, 80- tion, examination, analysis, 1-2, 20-21,
81, 83, 84, 87, 89-90, 92, 94, 97, 102, 24, 26, 28, 30, 40, 44, 46-48, 50, 53, 59,
102-103, 106, 122, 126, 132, 152-153, 60, 68, 70, 93, 100, 105, 147, 162, 166,
155, 186-187, 189, 190, 196, 198, 215, 175, 179, 187, 190, Chapter XI: 194-
232, 234, 250, 253, Pl. II, VII, XXIV 225; 227, 229-234, 241, 255, 257, Pl.
lamination, 50,59 I-XXXIX
lance, lancehead, 12, 55, 58, 60, 64, 66, meteorites, 5-6, 241
73-74, 185-186, 189-190, 196, 212, 215, meteoritic iron, 7, 9, 11, 13, 15, 20, 208,
217, 219-220, 227, 232, 234-235, Pl. 226-227, 234, 241
XXVIII, XXXVI Middle Ages see medieval
laterite, 241 Migration period, 46, 73, 75, 157, 197,
lattice of iron, 21 210, 216,-217, 251, Pl. III. XXV,
ledeburite, 50, 118, 194, 199, 241 XXXII-XXXIII
leucite, 115, 119 Mimung, the Sword, 75, 256
Limes Romanus, 147, 233-234, 250-251 minae, 6, 9, 11, 13, 15
lock, 139, 166-167, 187-188, 192, 238 mineral coal, 109-110, 112, 120, 181
locksmith, 71, 95, 100, 164-165, 192-193, mineralogical analyse, composition, 113-
231, 238 120, 139, 147, 149, 240
Lusatian culture, 81, 90, 251 mines, Chapter IX, 178-183
Lydians, 246 Minoan period, culture, 3, 14, 251
lynch pin, 58, 186-187, 192 Montelius, O., 3, 151, 245, 250, 252
monticellite, 119
murus Gallicus, muri Gallici, 192, 232,
magnetite, 110, 111, 112, 117, 119, 241, 251-252
243
Magyars, 162, 188, 251, 255
mail armour see armour nail, nails, nailmaking, 10, 14, 55, 58, 61,
malleability, malleable, 30, 35 71, 73, 75-76, 128, 144, 147, 167, 171-
mandrel, mandrels, 90, 167 172, 174-175, 177, 181-182, 186, 191,
martensite, 67, 69, 77, 81, 100, Chap- 196-197, 252
ter XI: 194-195, 198, 201, 205-206, nail hole in anvils, 76, 93, 95, 97, 98
208, 211, 216, 220, 222; 241-242, nail iron, 73-75, 97, 102, 146-147, 151, 165
Pl. I, XII, XIX, XXII-XXIII, XXVI, nail puller, 61, 192
XXVIII-XXXIII, XXXVI, XXXVIII- needle, 192
XXXIX Neumann bands, 76, Chapter XI: 194;,
medieval, Middle Ages, 14, 43-52, 56, 57, 241, Pl. XVII
58, 61, 67-68, 72, 75, 77, 78, 80, 82- nickel, 5, 8, 19-20, 37, 64, 208, 226, 241-
83, 84, 87, 89, 91-92, 93, 94-95, 97, 99, 242, Pl. XXII
101-104, 112, 123, 128, 130, 131-133, nitric substances, nitridation, 66, 70, 101,
160, 163, 164-187, 192-193, 196, 202- 197, 242
204, 206, 208, 211-212, 216, 222-224, nitrides, 70, 202, Pl. III
234, 237-239, Pl. XVIII, XXI, XXIII, Notitia dignitatum, 142, 230, 252
336 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
piling, 48, 59, 126-127, 161, 199, 205, Pl.
VIII-X, XX
ocel, steel, 19 plano-convex hearth bottom see PCB
Odyssey, 228, 249 plane, 193
Ochre Grave culture, 235 plating, 202, Pl. XIV
olivine, 119 Pliny the Elder, 231-232, 253
obeloi, obeliskoi, 37-39, 38, 191, 229 plough, ploughshare, 34-35, 36, 45, 122,
ornaments, 3-4, 6, 10, 14, 61, 72, 137, 184, 187-188, 196, 203, 208-210, 227-228, Pl.
186, 189, 196, 227-228, 232 III
Ostrogoths see Goths Plutarch, 39
oppidum, oppida, 23, 24, 33, 57, 63, 68, potash, 119
72, 77, 83, 84, 90, 100, 101, 103, 105, potlatch, 189
106, 116, 122, 125-126, 132, 136-137, Protogeometric period, culture in Greece,
140, 141, 153-156, 190, 201, 244, 251, 10, 17, 185, 253
255, Pl. VII, XII, XXIV Przeworsk culture, 186
Otomani culture, 8 punch, punching, 40, 58, 75-76, 90, 92,
186, 188, 200, 210, 212, 227, 234
punchmark, punchmarking, 39-40, 42,
padlock, 74, 122, 171, 192, 224-225, 231
165, 175, 189, 238
parzillu, iron, 4
pattern-welding, 2, 60, 63, 112, 213, 214-
222, 215, 221, , 231, 236, 239, 253, Pl.
XXXII-XXXV, XXXVI-XXXVIII quenching, 19, 68, 100-101, 195, 201, 205,
PCB, plano-convex cakes, cullots, 109, 12- 212-214, Pl. XXXII
120, 121, 125-126, 130, 125, 137, 146- quench-hardening, 67-69, 81, 100, 133,
151, 153-158, 160-165, 167-169, 171- 146, 194, 201, 205, 221, 227-230, 232,
175, 179, 181, 230 238, 242, 249
pearlite, pearlitic, 20, 24, 37, 40-41, 45-
46, 48, 50-51, 53, 60, 66, 105, 152,
Chapter XI: 194, 197-198, 201-203, rasp, rasps, 73-74, 101
205-212, 216-219, 223-224, Pl. I- razor, razors, 192
II, V-VIII, XI, XIII-XV, XVIII-XX, recrystallization, 66, 240
XXII, XXIV-XXVII, XXIX, XXXIII- recycling of iron, 23, 115, 122, 148, 155,
XXXIV, XXXVII, XXXIX 161, 181, 199
Pechenegi tribe, 164 Reinecke, P., 3, 25, 245, 253
Pen’kivka culture, 236, 252-253 ring mail see armour
Phaedon, tyrant of Athens, 38 Ringgriffmesser, 21, Pl. XII
Phoenicians, 114-115, 124, 138 rivet, riveting, 10, 58, 75, 90, 93, 173, 184,
phosphoric iron, 51, 60, 116, 154, 204, 212, 196, 222-223, 230, 236
216, 242 Pl. XVII, XIX Roman, Romans, 14, 17, 23, 34, 39-43, 43,
phosphorus, P-content in iron, 21, 26, 30, 59, 65, 67, 69, 72, 75, 76, 76-77, 78, 80,
37, 40, 48, 50, 64, 116, 200, Chapter XI: 82, 84, 85, 87-89, 91-94, 96, 98-100,
194-225; Pl. VIII, XVII, XIX, XXVIII 101-104, 105, 109-110, 111, 112, 116-
pick, picks, miner’s picks: 64, 178, 181- 117, 120, 121, 125, 128-129, 131-132,
183, 200, 207-209, 229, 238, Pl. X, 138-139, 141, 142, 145 151, 159, 160,
XXIII 188, 190, 192, 223-224, 230-231, 236,
pike, 192 247, 254
INDEX OF SUBJECTS 337
Romano-Barbarian period, 37, 58, 62, 66- 197, 202, 209, 220, 235-237, 244, 250-
67, 72-73, 79, 88, 91, 99, 103,, 105, 251, 253-255, 257, Pl. III, XXVIII
113, 116-117, 132, 135, 157, 159, 172, smithies, smithy, 1, 48, 71, 109, 111, 112,
190, 192, 198, 201-202, 216, 253-254, 114-116, 118, 120, 121, 122, 125, 127,
Pl. VII, IX, XIII 129, 131-134, Chapter IX:135-183; 188,
221, 224, 229, 232, 238, 244, 247, 253,
256
sabre, 192, 207, 244, 246 smithing slag, see also PCB, 1, 10, 72,
sacrifice (votive) deposit, place, 75, 189, Chapter VII: 112-121, 120, 121, 125,
249-250, 254, 256 134, Chapter IX: 135-138, 151, 153,
sallet, 223 164, 242, 245, 248
Saltovo-Mayatsk culture, 236, 250, 254 smithing, smith’s, blacksmith’s tools, 12,
Sarmatians, Sauromatae, 235, 251, 254- 33, Chapter VI: 71-108, 76, 88, 100,
255 101, 135, 145, 155, 166, 175,184, 189,
sax, 73 190, 192, 232, 237, 251, 255
scramasax, 73, 215, 217-218, 236, 254, Pl. soldering, 75, 115, 120, 192, 225, 238
XXXV Sophocles, 229, 255
saw, 201, 228, Pl. VII sorbite, 69, 195, 208, 210, 242 Pl. XIX,
Saxon see Anglo-Saxon XXV-XXVII, XXXVIII
scabbard, 189, 197, 232, 235-236 spade-shaped bars, 43, 44, 237
spatha, spathae, 215, 236, 255
Schedula diversarum artium see
spear, spearhead, 12, 14, 64, 74, 162, 177,
Theophilus Presbyter
187-188, 202, Pl. III, XI
scissors, 58
spit, spits, 36, 37, 38, 39, 186, 229
scrap iron, metal, 23, 48, 115, 122, 135,
spoils, 4, 12, 139, 234, 252
137, 146, 151, 155, 167, 199, 231
spur, spurs, 122, 187
scythe, scythes, 69, 96. 187-189, 192
steel, 1, Chapter III: 18-22;37, 45, 51,
scythe-shaped bars, 44, 237
53-54, 58, 60, 63, 66, 67, 68, 69-71,
Scythian, Scythians, 10, 17, 58, 62, 84, 77, 81, 99-100, 110, 114, 117, 119-120,
153, 233, 235, 246, 254 123, 131, 133, 181, 187, 194, 196,198,
sheet, 46, 55, 122, 144, 155, 166, 187, 192, 201-222, 228, 230, 232-235, 237-238,
199, 239 241-242, 245-246, 249, 256, Pl. IV,
sheet shears, 58, 72-75, 104, 105, 145, 189, VI-VIII, XIV-XVI, XVIII, XIX-XXV,
230 XXVII-XXVIII, XXX-XXXI, XXXIV-
shekel, 6, 7, 9, 11, 13, 15, 226 XXXV, XXXVIII-XXXIX
shield, 189 stirrup, 188
shield boss, 73, 187, 190, 230 stómoma, steel, 19
shieldmaker, 187 stone anvil see anvil stone
shovel, 192 strike-a-fire, strike-a-light, 62, 63, 177,
sickle, sickles, 14, 37, 38, 69, 74-77, 96, 180, 187, 192, Pl. XXVIII
162, 175, 187-188, 190, 209, 229, 234 stripe welding, striped damast, 60, 212,
sidareoi, 39, 191, 229 213, 214
sidéros, iron, 19 structural iron/steel, 14, 21, 137, 139, 147,
slag inclusions, 40, 46, 50, 59, 205, 210, 184, 197, 206, 229, 238, Pl. V
241, Pl. XXV Submycenaean period, 17
slaves, 72 swages, swaging, 54, 63, 64, 74, 93, 98,
Slavs, Slavic, 46, 144, 161-162, 169, 186, 101, 230
338 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
sword, swords, 5, 12, 14, 34-35, 36, 58, Ulfberht, the swordsmith, 236, 256
60, 63, 64, 66, 70, 73-74, 120, 124, 174,
184, 185, 186-187, 189, 190, 191, 196-
197, 203,-204, 209, 215, 218, 227-228, Varni tribe, 215, 234
232, 234-236, 246, 254, Pl. II, VIII, Vendel period, 43, 45, 234
XIV, XVI, XVII, XXXII-XXXIV Venedoi tribe, 254
swordsmith, 14, 178, 190, 204, 216, 229, Veneti tribe, 249, 251, 256
231, 238, 244 vicus, vici, 143, 146, 192, 205, 248, 256
Viking period, 43, 50, 72-73, 75, 79, 83,
88. 102-104, 105, 121, 132, 134, 160,
Tabula Peutingeriana, 146 170, 172-173, 211, 219, 234, 248, 256,
Tacitus, P. C., 192, 255 Pl. XX, XXVIII-XXIX, XXXVI
taenite, 19, 242 villa rustica, 40, 109, 121, 127, 129, 130,
taleae ferreae, 32-37, 232 134, 139, 230, 141, 148-151, 230, 256
talents, 7, 9, 11, 12, 13, 15 Visigoths see Goths
Tchernyakhov culture, 247, 255 votive deposit see sacrifice deposit
temenos, 26, 255 Vulcanus, 132, 247
tempering, 19, 68-69, 195, 201, 217, 220,
227-229, 238, 242, Pl. XXXIX
terrestrial iron, 20 waste, smithing, metallurgical, 1, 3, 8,
Theophilus Presbyter, 67, 84, 100-101, 10, Chapter VII: 109-122; 134-135, 140,
177, 255 146-147, 154, 158, 160-161, 166, 177-
Theodorich the Great, 215, 236, 247 179
Thracians, 246 water tanks, throughs, barrels, 130, Chap-
tiles, 126, 149, 153 ter VIII: 133; 166, 170, 173, 182
tilling, tillage Pl. III, XV weapons, weaponry, 3, 10, 14, 52-53, 55,
Tischler, O., 250 58, 64, 73-75, 164, 168, 175, 184-190,
tongs, 58, 71-75, 81, 84-90, 85-90, 97, 105, 192, 196-198, 203, 212, 230-232, 235-
137-138, 144, 151-152, 156, 164, 172, 237, 246, 248, 250-251, 254, 257, Pl.
175, 177, 181, 186, 230 III, XXXV
tools, toolmakers, toolmaking, 3, 10, 12, welding, welds, weldable, 2, 18, 20-21,
14, 22, 33, 37, 52-53, 58, 66, 68, Chap- 30, 35, 50, 52, 54-55, 58-60, 65, 90,
ter VI: 71-102; 122, 135, 139, 155, 162, 93, 100, 118, 127-128, 143-144, 151,
166, 173, 175, 179, 184, 186-187, 189, 161, 181, 192, 196, 200, 202, Pl. V,
192, 198-200, 203, 208, 211-212, 231- IX, XII, XIV-XVIII, XXIV-XXVIII,
232, 236, 238-239, 245, 250, Pl. IV, X XXX-XXXVIII
tributes, taxes, 12, 50, 252 welding seam, 21, 30, 45, 60, 199, 202-203,
tripod, 37, 39, 74, 188-189, 191 205-206, 208, 211, 226, 229, 232, 234-
trompe, Bottichgebläse, 127, 129, 132, 237, 241-242, 254, 257, Pl. XII, XV-
150, 159 XVI, XXI, XXVI, XXXI, XXXVIII
troostite, 69, 100, 194, 211-212, 220, Pl. wheel tyre, tyres, 14, 58, 64, 125, 147,187,
XXIX, XXXII, XXXIII, XXXVI 193
tuyere, 109, 113, 121, 125-126, 132-133, wheel naves, 59, 63, 64, 139, 186-187,192,
142, 146, 149, 155-156, 161-162, 164, 235, 245, 251
169, 171-172, 178, 200, 242 wheelwright, cartwright, 64, 125
tyre see wheel tyre whetstone, 153, 162, 172
tweezers, 81, 188 Weyland the Smith, 70, 244-245, 256
INDEX OF SUBJECTS 339
Widmannstätten structure, 6, 20-21, 81, wüstite, 110, 111, 112, 114, 119, 154, 240-
194, 198-199, 201-203, 206, 208, 210, 241, 243
241-242, Pl. I, IV, VIII, XX
wire, 54-55, 76, 88, 83, 155, 164, 171, Pl.
XXXVI Xenophon, 245, 257
wire-drawing, 73-75, 97, 98, 101-102, 102,
104-105, 211-212, 230, Pl. XXIX,
XXXVI
wire-drawing iron, die, hole, 211-212 Zarubincy culture, 235, 257
wrought iron, 42, 53, 59-60, 118, 196-198, železo, żelazo, iron, 19
243, Pl. X
340 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS
PLATES 341

Pl. I: Metallographic structures. 1 Ferrite, 100x; 2 pearlite with ferritic network, 100x;
3 Widmannstätten texture of pearlite (dark) and ferrite, 200x; 4 martensite, 1100x; 5
bainite, 200x; 6 sorbite, 1000x.
342 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. II: All-iron sword blade. Třebohostice, Bohemia, sample 178, La Tène period. 1 Blade
with specimen positions marked; 2 transversal section composed of microphotographs,
23x; 3 microhardness mHV 30g; 4 wedge-shaped weld, 100x; 5 - 6 ferrite with interstitial
pearlite, 150x; 7 thin adhering layer of a copper-based metal (left), 100x. Etched with 2
% Nital.
PLATES 343

Pl. III: All-iron weapon and tilling tool. 1 - 4 Klučov, Bohemia, spear head sample 1,
Migration period. 1 Sampling; 2 scheme of the transversal section; 3 - 4 various grain sizes
of ferrite, 220x. 5 - 7 Klučov, sample 45, Slavic. 5 Symmetrical ploughshare, sampling;
6 cross-section scheme with marked nitride needles, 7x; 7 coarse and fine ferrite, 200x.
Etched with 2 % Nital.
344 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. IV: Application of mild steel in tool manufacture. Kilberg, Ireland, 3rd century
BC/5th century AD shaft-hole axehead, sample 19. 1 Sampling; 2 cross-section of the
cutting-edge and microhardness; 3 Widmannstätten texture, 100x. Etched with 4 % Nital.
After Scott.
PLATES 345

Pl. V: Forming, folding and welding of structural iron. 1 - 3 Persepolis, Iran, heteroge-
neously carburized iron clamp, 6th century BC. 1 Cross-section, 2 - 3 ferrite and pearlite,
pearlite, 200x. After Pleiner. 4 Athens, Parthenon, 5th century BC. Clamp of pearlitic
steel and iron bands welded into an H-shape. Scale in cm. After Varoufakis.
346 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. VI: Early all-steel blade. Radzovce, Slovakia, curved knife of the 8th century BC. 1
Sampling; 2 Cross-section scheme and microhardness mHV 30g; 3 - 4 pearlite with an
excess cementite network, 150x and 75x; Etched with 2 % Nital.
PLATES 347

Pl. VII: All-steel and additionally carburized artefacts. 1 - 4 Abrahám, S Slovakia,


Romano-Barbarian strike-a-fire sample 567, 1st century AD. 1 Engraved artefact and
sampling; 2 schematic section and microhardness mHV 30g; 3 - 4 pearlite and excess
cementite, 100x. 5 - 7 Stradonice Oppidum, Bohemia, La Tène period. 5 foxtail saw
fragment; 6 scheme of the cross-section, 5x; 7 fine ferrite-and-pearlite, 200x. Etched with
2 % Nital.
348 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. VIII: Piling of mild steel bands. Singen, SW Germany, sword blade of the 9th/8th
century BC. 1 Sampling; 2 piled texture along the cutting-edge in specimen 4; 3 pearlite
and ferrite network in a carbon-richer zone of specimen 2; 4 specimen 3 - bands of P-rich
(dark) and P-poor (light) iron, etched after Oberhoffer; 5 Widmannstätten with globular
pearlite; 6 blade cross-section B with the main split weld marked; 7 the same, ferrite and
ferrite-and-pearlite structures. After Boll et. al.
PLATES 349

Pl. IX: Piling and folding of heterogeneously carburized iron. Kostolná, S Slovakia, sample
583, Romano-Barbarian knife of the 1st century AD. 1 Sampling; 2 cross-section after
Oberhoffer etching; 3 the samme after 5 % Nital etching, 14x; 4 structures of a welded-in
wire; 5 fine ferrite along the main weld, 4 and 5 10x; 6 the weld in the cutting-edge; 2 to
5 etched with Nital.
350 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. X: Piling of carbon-poor iron in toolmaking. Čáslav-Hrádek, E Bohemia, stone-


cutter’s pick of the 11th century AD. 1 Double-edged pick and sampling; 2 schemes of
both cutting-edges, 4x; 3 - 5 ferrite, impure wrought iron, 180x. Etched with 2 % Nital.
PLATES 351

Pl. XI: Secondary carburization of blades. 1 - 4 Hallstatt period knife sample 62 from a
princely grave at Lovosice, 6th century BC. 1 Sampling; 2 blade cross-section (carburized
metal dotted); 3 ferrite in the blade body, 200x; 4 penetration of pearlite into the cutting-
edge, 200x. Etched with 2 % Nital. After Pleiner. 5 - 7 Hillsborough, Ireland, early
medieval spearhead; 5 Sampling; 6 macrophotograph of the carburized edge; 7 carbon
gradient: pearlite penetrates into the iron body, 100x. Etched with 4 % Nital. After
Scott.
352 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XII: Piling of carburized bands. Ringgriffmesser’ type knife 503 from the Celtic
oppidum of Hostýn, Moravia. 1 Sampling; 2 cross-section etched after Oberhoffer; 3
etched after Heyn, 5.5x; 4 cross-section scheme with microhardness curve, mHV 30g; 5
blade back: welding seam, black zone - martensite, 70x; 6 middle of the blade: welding
seams, 75x; 7 - 8 cutting-edge: carburized zone with martensite, both 120x. 4 - 8 etched
with 2 % Nital.
PLATES 353

Pl. XIII: Secondary carburization of a knife blade. Sládkovičovo-Tornal’a, S Slovakia,


Early Romano-Barbarian knife 555. 1 The object; 2 - 3 cross-section etched after Ober-
hoffer, 6. 5x; 4 cutting-edge, composed of microphotographs, 18x - increasing amount of
pearlite (dark); 5 - 7 ferrite and pearlite (dark), 100x. 5 - 7 etched with 2 % Nital.
354 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XIV: Iron and steel plating. Holubice 606, Moravia, Celtic sword. 1 Sampling; 2
etched cross-sections, 5x (inversed, carbon-rich parts light); 3 microhardness curve mHV
30g, solid line: pearlitic areas; 4 left cutting-edge: steel parts dark, 26x; 5 - 6 dark pearlite
and lighter ferrite, welds, 100x. 5 - 6 etched with 2 % Nital.
PLATES 355

Pl. XV: Welding-on steel as applied on tillage implements. 1 - 4 Ivanovice, Moravia,


ploughshare of the 9th century AD. 1 Sampling; 2 scheme of the edge (scarf-welded steel
dotted); 3 pearlitic (dark) and ferritic plates in the edge, 50x; 4 pearlite with ferrite
network, 200x. After Pleiner. 5 - 8 White Fort, Ireland, early medieval coulter. 5
Sampling; 6 welding seam between low- and high-carbon areas; 7 ferrite, welds, 100x; 8
spheroidized pearlite, 500x. 6 - 7 etched with 4 % Nital, 8 etched with Picral. After Scott.
356 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XVI: Welded-on steel shell on a weapon blade. Cyprus, Museum collection, dirk or
dagger from the 11th century BC. 1 Sampling; 2 scheme of a half-section; 3 welding seam
joining the ferritic core and the steel shells, carbon diffusion across the weld, 300x. After
Lang.
PLATES 357

Pl. XVII: Welding-on of phosphorus-rich shells. Jenišův Újezd, NW Bohemia, Celtic


sword sample 586. 1 Sampling; 2 cross-section, etched after Oberhoffer, phosphoric iron
side shells dark, 10x; 3 microhardness mHV 30g - side shells hatched and dotted; 4 cross-
section composed of microphorographs, 19x; 5 fine ferrite in the core, coarse ferrite in the
shells, 100x; 6 Neumann bands in a ferrite grain from the right shell, 200x. 4 - 6 etched
with 2 % Nital.
358 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XVIII: Application of the three layer sandwich in cutlery. 1 - 2 Helgö, Sweden, knife
of the 6th/7th century AD. 1 Cross-section composed of microphotographs - dark steel
core and weld with segregated ‘white lines’, 100x; 2 tang: fine pearlitic S-shaped stripe
marking the weld of two iron bands, 150x. Etched with Nital. After Modin and Pleiner.
3 Slobodka, Belarus, cross-section of an early medieval knife; 4 Kletsk, Belarus, knife,
sandwich with additional side steel shells. 3 - 4 after Gurin.
PLATES 359

Pl. XIX: Application of the iron-and-steel three-layer system. 1 - 4 Glazov, Russia, knife
from the 10th century AD. 1 Sampling; 2 cross-section, dark steel central strip; 3 etched
after Oberhoffer, phosphoric iron side shells; 4 clean weld between ferritic and martensitic
cutting-edge. After Kolchin. 5 - 7 Ivanovice, Moravia, plougshare sample 137, 9th century
AD. 5 Sampling; 6 right edge section, steel is dark, 4.5x; 7 sorbitic and pearlitic central
plate, 50x. 6 - 7 etched with 2 % Nital.
360 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XX: Welding-in an iron-and-steel pile. Haithabu, Jutland, Viking period axehead. 1
Object No 1; 2 upper part, completely sectioned, 1 : 1x; 4 scheme of the section (symbols
give approximate C-contents). 5 macrophotograph of the cutting-blade (steel is dark),
6x; 6 ferrite and pearlite in the edge, 200x; 7 pearlite and ferrite in a Widmannstätten
texture, 200x. After Thomsen.
PLATES 361

Pl. XXI: Inserting steel cutting edges into axeheads and knife blades. 1 - 4 Ballynahinch,
Ireland, sample 51, early medieval. 1 Upper side of the axehead; 2 blade cross-section
with inserted dark steel plate; 3 welding seam joining the iron and steel components, 50x;
4 reconstruction of the manufacturing of the axehead. After Scott. 5 Menka, Belarus,
cross-section of a knife blade. 5 after Gurin.
362 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XXII: Folded blade inserted into the socket. Kjula, Sweden, axehead (presumably
the 1st half of the 1st millennium AD). 1 Axehead; 2 scheme of the cross-section; 3
macrophotograph of the folded blade, 7x; 4 Ni-rich martensitic streaks in the ferritic-and-
pearlitic matrix, 35x; 5 - 6 details, 400x and 950x. After Hermelin et al..
PLATES 363

Pl. XXIII: Inserting a steel point into a medieval mining pick. Pampailly, France. 1
Scheme of the sectioned tool; 2 martensite and bainite of the steel point. After Benoit.
364 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XXIV: Scarf welding-on of steel cutting-edge on an axehead. Manching oppidum,


Bavaria, La Tène period. 1 Sampling of a socketed axehead; 2 sample scheme; 3 recon-
structed manufacture; 4 curved welds in the interstitial band along the cutting-edge; 5
steel plate in the cutting-edge, 50x; 6 - 7 ferrite and pearlite in the same place, 200x and
100x. Etched with 2 % Nital.
PLATES 365

Pl. XXV: Scarf welding-on of steel in cutlery. 1 - 6 Březno, NW Bohemia, Migration period
knife, sample 619. 1 Sampling; 2 distribution of slag inclusions in the cross-section, 10x; 3
scheme of the cross-section, microhardness curve added (mHV 30g); 5 blade back: ferrite
and pearlite, 75x; 6 cutting-edge: fine pearlite/sorbite, 70x. Etched with 2 % Nital. After
Pleiner. 7 Navasiolki, Belarus, medieval knife cross-section, after Gurin.
366 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XXVI: Scarf welding-on of carburized cutting-edge. Mikulčice, Moravia, battleaxe


sample 175, 9th century AD. 1 Sampling; 2 cutting-edge scheme, 4x; 3 - 4 welded layers
in the cutting-edge, sorbite, martensite, ferrite, 100x; 5 ferrite and pearlite in the middle
of the blade, 200x; 6 - 7 ferrite and pearlite in the low-C parts of the cutting-edge, 200x.
Etched with 2 % Nital.
PLATES 367

Pl. XXVII: Scarf- and butt-welding of steel in cutlery. Budeč hillfort, Bohemia, 10th
century AD. 1 Knife fragment; 2 cross-section, etched after Oberhoffer, 6x; 3 etching
after Heyn, microhardnes curves (mHV 30g) added; 4 cutting-edge part composed of
microphotographs, 20x; 5 cutting-edge: pearlite-and-ferrite, 150x; 6 cutting-edge: sorbite,
200x. Etched with 2% Nital.
368 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XXVIII: Butt-welding of steel. 1 - 2 Newbury area, England, barbed arrowhead


from the Viking period. 1 Arrowhead, 2 martensite in the barbs. After Brown. 3 -
4 Lutomiersk, Poland, early medieval Slavic lancehead. 3 Lance; 4 part of transversal
section: dark steel, grey iron, lustre P-rich interlayer. After Piaskowski. 5 - 8 Nitra-
Lupka, Slovakia, Slavic hillfort, strike-a-fire 107 from the 9th century AD. 5 Sampling;
6 cross-section, dark steel striking edge, 4x; 7 the same, 50x; 8 edge: martensite, 500x,
etched with 2 % Nital. After Pleiner.
PLATES 369

Pl. XXIX: Heat treatment applied on a wire-drawing iron. Haithabu, Jutland, Viking
period, artefact 1. 1 Three views of the tool, left the running side; 2 macrophotograph
of a cross-section, 7x; 3 martensite, 500x; 4 martensite and dark troostite, 100x; 5 ferrite
and pearlite, 500x. Etched with Pikrin acid. After Naumann.
370 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XXX: Butt-welding of steel cutting-edges in cutlery. 1 - 4 Čáslav-Hrádek, Bohemia,


early medieval knife sample 23. 1 Sampling; 2 cross-section scheme, 11x; 3 blade back:
ferrite, 200x; 4 cutting-edge: martensite and traces of ferrite, 200x. Etched with 2 % Nital.
After Pleiner. 5 Lemeshevichi, Belarus, cross-sectioned medieval knife. After Gurin.
PLATES 371

Pl. XXXI: Striped damast as applied on medieval knives. Hradištko-Sekanka, Bohemia,


13th century. 1 - 5 Knife sample 251. 1 Sampling; 2 Cross-section 9x, etched after
Heyn and microhardness curve (mHV 30g); 3 blade back: steel interlayer, welding seam,
carbon diffusion, 100x; 4 the same place, 200x; 5 cutting-edge, martensite, 100x. 6 - 7
Knife sample 258, steel cutting-edge and serrated weld; 7 cross-section scheme. Etched
with 2 % Nital.
372 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XXXII: Pattern-welded panels in a sword blade. Bešeňov, Slovakia, sample 606,
Migration period. 1 Split soft cutting-edge, part of a cross-section composed of micropho-
tographs, 50x; 2 hard cutting-edge, quenched: martensite and troostite, 50x; 3 part of the
pattern-welded panel in the middle of the blade, 50x; 4 cross-section scheme, 5x. Etched
with 2 % Nital. After Pleiner.
PLATES 373

Pl. XXXIII: Bešeňov (continued). 1 Cross-section of the sword blade, etched after Ober-
hoffer, microhardness curve mHV 30g added, 3x; pattern-welded panel, alternating stripes
of ferrite-and-pearlite and insufficiently etched ferrite, 100x; 3 cutting-edge: martensite
and dark troostite, 100x. Etched with 2 % Nital. After Pleiner.
374 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XXXIV: Manufacture of an early medieval princely sword. Kolı́n, Bohemia, 10th
century AD, sample 10. 1 Sampling; 2 surface polishing, traces of a pattern-welded panel;
3 cutting-edge cross-section at A, steel appears as dark, 12x; 4 pearlite and ferrite in
the core, 200x; 5 pearlite and interstitial pearlite in the cutting-edge, 200x. Etched with
potassium iodide. After Pleiner et al..
PLATES 375

Pl. XXXV: Pattern-welding on weapons and knives. 1 - 3 Saxon scramasax from Dorset
Museum. 1 Radiograph; 2 Blade scheme with pattern-welded panels; 3 blade back: iron
and steel bands, 75x. After Brewer. 4 Winchester, England, 13th century sectioned knife
with pattern-welded back. After Tylecote. 5 Gdańsk, Poland, early medieval knife. After
Piaskowski.
376 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XXXVI: Manufacture of an early medieval display lancehead. Haithabu, Jutland,


Viking period. 1 Lance with welded-on socket trunnions; 2 schematic view with partly
pattern-welded core and serrated welds at the edges; 3 cross-section of the blade, 5x; 4
serrated welds, 1 : 1; 5 macrophotograph of the pattern-welded part, 5x; 6 serrating with
an imprinted twisted wire, radiograph, 2x; 7 - 8 tempered structure of martensite and
troostite, 500x and 75x. After Thomsen.
PLATES 377

Pl. XXXVII: Pattern-welded knives. 1 - 8 Bukhak, Ukraine, 13th century AD. 1 Sampling;
2 schematized cross-section, 9x; 3 - 8 alternating stripes of ferrite and ferrite-and-pearlite,
70x. After Voznesenskaya. 9 Etched cross-section of a medieval knife from Vitebsk,
Belarus. After Gurin.
378 IRON IN ARCHAEOLOGY: EARLY EUROPEAN BLACKSMITHS

Pl. XXXVIII: Pattern-welded cutlery. Hradištko-Sekanka, Bohemia, sample 249, 13th


century AD. 1 Polished surface of the blade, sampling; 2 cross-section etched after Ober-
hoffer and Heyn, microhardness curve added, mHV 30g. 3 blade back: ferrite and in-
terstitial sorbite, welding seam, 100x; 4 pattern-welded middle of the blade: alternating
hard (dark) and softer steel bands, 100x; 5 cutting-edge: ferrite and martensite, welding
seam, martensite, 100x. Etched with 2 % Nital.
PLATES 379

Pl. XXXIX: Steel in the manufacture of medieval plate armour. 1 - 3 Rennhut type
helmet from the Kunsthistorisches Museum in Vienna B 129, about AD 1490. 1 Helmet;
2 tempered martensite with traces of ferrite, 50x, etched with Nital; 3 the same place,
200x. 4 - 5 Breastplate covered with red velvet, Bayerisches Nationamuseum, Munich,
W 195. 3 Breastplate; 4 very fine pearlite mixed with martensite, 70x. Etched with
Nital/Picral. Courtesy of A. R. Williams.

You might also like